Next Article in Journal
Radiolabeling of Platelets with 99mTc-HYNIC-Duramycin for In Vivo Imaging Studies
Previous Article in Journal
Genome-Wide Identification and Comprehensive Analysis of the GASA Gene Family in Peanuts (Arachis hypogaea L.) under Abiotic Stress
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Regulation of Peptidase Activity beyond the Active Site in Human Health and Disease

Faculty of Chemistry and Chemical Technology, Department of Chemistry and Biochemistry, University of Ljubljana, Večna pot 113, SI-1000 Ljubljana, Slovenia
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(23), 17120; https://doi.org/10.3390/ijms242317120
Submission received: 17 November 2023 / Revised: 1 December 2023 / Accepted: 2 December 2023 / Published: 4 December 2023
(This article belongs to the Section Biochemistry)

Abstract

:
This comprehensive review addresses the intricate and multifaceted regulation of peptidase activity in human health and disease, providing a comprehensive investigation that extends well beyond the boundaries of the active site. Our review focuses on multiple mechanisms and highlights the important role of exosites, allosteric sites, and processes involved in zymogen activation. These mechanisms play a central role in shaping the complex world of peptidase function and are promising potential targets for the development of innovative drugs and therapeutic interventions. The review also briefly discusses the influence of glycosaminoglycans and non-inhibitory binding proteins on enzyme activities. Understanding their role may be a crucial factor in the development of therapeutic strategies. By elucidating the intricate web of regulatory mechanisms that control peptidase activity, this review deepens our understanding in this field and provides a roadmap for various strategies to influence and modulate peptidase activity.

1. Introduction

Peptidases, enzymes that cleave the peptide bond of (poly)peptides, are ubiquitous in nature. They are found in all three domains of life as well as viruses, and they make up a significant proportion of the enzymes found in living organisms. The MEROPS database (available online at https://www.ebi.ac.uk/merops/, accessed on 7 March 2023) catalogues over 4000 peptidases, of which approximately 700 are found in humans [1]. This wide range comprises peptidases of seven different catalytic types, i.e., serine, cysteine, aspartic, glutamic, asparagine and metallopeptidases, underscoring the multifaceted role that these enzymes play in regulating important physiological processes such as blood coagulation and apoptosis as well as nonspecific protein turnover. Given their diverse roles, a tight regulation of peptidase activity is required, and their dysregulation has been linked to the development or progression of numerous diseases, including autoimmune diseases, cancer, cardiovascular disease, and bacterial or viral infections. As of 2022, 25 drugs targeting peptidases were in clinical use and nearly 100 others were in various stages of clinical testing [2].
The regulation of proteolytic enzymes occurs at several levels, starting with transcription. Many proteolytic enzymes are expressed as inactive zymogens that require specific conditions for activation. After activation, peptidase activity can be modulated by ligands that bind to the active site or other areas of the enzyme surface, such as exosites (i.e., additional substrate binding sites distant from the active site) and allosteric regulatory sites. Several mechanisms can be involved in the regulation of one and the same peptidase, as illustrated on the example of thrombin in Figure 1 [3]. The modulators can be organic molecules, peptides, or entire proteins of endogenous or synthetic origin, and they can influence not only the enzyme activity but also the specificity of the peptidase. The vast majority of modulators targeting peptidases for medicinal and/or research purposes are inhibitors that bind to the active site in substrate-like manner, thereby abolishing enzyme activity. While this is a proven and effective strategy in some cases, like the gliptin class of dipeptidyl-peptidase 4 inhibitors for the treatment of type 2 diabetes [4], it can also cause significant on or off-target side effects that can ultimately prohibit the use of the inhibitor in the clinic. For example, the cathepsin K inhibitor odanacatib (Merck & Co., Rahway, NJ, USA) was terminated after completing phase 3 clinical trials for the treatment of post-menopausal osteoporosis due to rare but significant side effects, despite continued demonstration of its efficacy [5]. Exploring alternative ways of regulating peptidase activity in such cases could prove beneficial in comparison to active site-directed inhibitors.
The aim of this review is to evaluate a range of peptidase–modifier interactions by examining and comparing some recently discovered and common mechanisms, such as the activation of zymogens, binding of glycosaminoglycans, regulation via exosites and allosteric sites, and non-inhibitory binding proteins. We concentrate on interactions relevant to human physiology and pathology in order to provide an expert opinion on whether these strategies can be used to regulate peptidase activity in vivo.

2. Regulating the Activation and Maturation of Zymogens

One of the most fundamental and widespread mechanisms for protection against excessive proteolysis is the synthesis and secretion of peptidases in the form of inactive precursors called zymogens. These inactive peptidases are stored until activation signals are received, allowing for rapid recruitment without the need for prior transcription. Zymogens contain inhibitory (pro)peptides that prevent substrate binding and proteolytic activity. Therefore, the inhibition of activation is a viable strategy for preventing peptidase activity in general.
One of the earliest studies of zymogen activation, which incidentally is quite different from other peptidase families, is the chymotrypsin-like serine peptidases. Here, zymogen inactivity results from incomplete folding of the peptidase (Figure 2a) rather than physical closure of the active site by the proregion. Importantly, the S1 substrate-binding site and the oxyanion hole are not formed in the inactive state. Once the N-terminal hexapeptide, known as the activation peptide, is cleaved, the newly formed N-terminal residue forms a salt bridge that triggers a conformational change leading to the formation of the active site [6,7,8]. The activation peptide can be cleaved by autocatalysis or by other peptidases [9,10,11,12]. An example of clinically relevant activators of chymtrypsin-like serine peptidases is dipeptidyl peptidase I (DPPI, also known as cathepsin C), which activates most of the immune system serine peptidases, including granzymes [13], neutrophil serine peptidases [14], and mast cell chymase [15]. As an important catalyst for the activation of peptidases in immune cells, DPPI is a promising target for the treatment of inflammatory diseases characterized by immune system overactivity. To date, numerous DPPI inhibitors have been developed and tested (see [16] for a recent review). The most advanced candidate for clinical use, brensocatib (Insmed, Inc., Bridgewater, NJ, USA), is currently being tested in a phase 3 clinical trial for the treatment of non-cystic fibrosis bronchiectasis, following the successful completion of a phase 2 trial [17].
DPPI belongs to the papain-like peptidase family, which includes members collectively known as cysteine cathepsins in animals. Most cysteine cathepsins (but not DPPI) are capable of autocatalytic activation. Their proregion is reversibly linked to the fully formed active site and physically blocks the cleft, preventing substrate binding. The activation mechanism of cysteine cathepsins has been extensively studied in procathepsin B, which is a cysteine peptidase involved in intracellular catabolism, the regulation of angiogenesis [18], neurogenesis [19], and also tumor growth and invasion [20,21]. Binding of the proregion is pH dependent [22], and at acidic pH, the electrostatic interactions between the residues in the proregion and the catalytic cleft are disrupted, leading to the release of the proregion and the formation of an intermediate structure, the active zymogen. This active zymogen can then activate adjacent proenzyme molecules by proteolytic cleavage of their proregion [23] in a bimolecular process [24]. Cathepsin K is another well-studied cysteine cathepsin involved in bone remodeling that is capable of autocatalytic processing in acidic bone remodeling lacunae [25]. Other lysosomal cysteine cathepsins such as L [26,27,28], S [29] and F [30] are also capable of autocatalytic activation.
Some cysteine cathepsins cannot undergo autocatalytic processing, such as human cathepsin X, which has a short proregion covalently linked to the active site via an unbound cysteine residue in the proregion and a cysteine residue in the catalytic diad [31]. As a pure exopeptidase, cathepsin X is processed by other cysteine peptidases, such as cathepsin L [32]. Similarly, DPPI cannot be autocatalytically activated. DPPI has a unique additional exclusion domain that converts its typical papain-like endopeptidase activity to a dipeptidylpeptidase [33,34], and it is activated by either cathepsin L or S [35]. Procathepsin H is another papain-like cysteine peptidase that stands out from the family, as it is cleaved at three different sites upon activation. The resulting active enzyme contains both a catalytically active domain and a minichain derived from the propeptide that remains covalently bound to the enzyme. This minichain occludes the active site and enables the aminopeptidase activity of the peptidase [36]. Cathepsin L is believed to activate procathepsin H [37].
Similar binding of the unpaired cysteine residue as in cathepsin X can be observed in zymogens of matrix metalloproteases (MMPs; Figure 2b), where this highly conserved residue interacts as a fourth ligand for the zinc ion in the active site. This binding, termed the cysteine switch [38], excludes water from the coordination sphere of zinc (Figure 2b). To activate MMPs, this bond must be disrupted, and the thiol group of the cysteine must be replaced by a water molecule. This is accomplished by stepwise proteolytic cleavage followed by conformational changes [39]. Peptidases such as kallikreins and the MMPs themselves are involved in cleavage [40]. Cytotoxic oxidants such as hypochlorous acid, a product of myeloperoxidase, can also activate the cysteine switch in metalloproteases [41,42,43].
Aspartic peptidases, including pepsinogen and related peptidases, follow a similar activation process as cysteine peptidases. The proregion of pepsinogen is bound to the catalytic cleft by reversible electrostatic interactions [44]. Lowering the pH weakens these interactions and causes a conformational change [45]. This structural change creates a transition state that leads to the formation of the active enzyme through further proteolytic cleavage. Cleavage can occur either by autoproteolysis, as in the case of pepsinogen [46], or by other peptidases, as in the case of prorenin [47]. In the case of procathepsin D, which is also an aspartic peptidase, both autocatalytic processing and processing by other peptidases are possible [48,49,50,51]. However, the autocatalytic processing of procathepsin D requires the presence of sulfated glycosaminoglycans (GAGs) [52].
Targeting zymogen activation may be a promising therapeutic strategy, but it is not always effective. For example, analyses of the inhibitory capacity and specificity of the proregion of individual members of the human cysteine cathepsins [53,54,55,56] have shown that for closely related family members such as cathepsins K, S, and L, selectivity is insufficient for specific protein targeting [56]. Targeting the activation of short-lived peptidases that would otherwise require high local concentrations of the inhibitor is likely to be more successful. Such peptidases are peptidases of the complement system, which is an important component of innate immunity. This strategy is employed in nature by tick Rhipicephalus Pulchellus, which has evolved a mechanism to inhibit host complement peptidases to evade the immune system. The saliva of the tick contains CirpT: short peptides that bind to the complement peptidase zymogen and prevent its activation (Figure 3) [57]. A similar effect is achieved with a small molecule that inhibits the activation of the matriptase zymogen [58]. Matriptase is a transmembrane serine peptidase involved in extracellular matrix degradation and epithelial migration and consequently involved in processes of cancer invasion and metastasis [59,60]. Inhibition of its activation is therefore a potentially clinically important process in the treatment of cancer.
A potentially clinically useful regulator of zymogen activation is also a monoclonal antibody that binds to the proenzyme of urokinase-type plasminogen activator (uPA), which is an enzyme that converts plasminogen to plasmin [61] and is involved in tumor progression and metastasis. Binding of the antibody delays cleavage of the proregion and the subsequent activation of uPA [62]. Recently, a small molecule (JNJ0966) was found to inhibit the activation of metalloproteinase-9 (MMP-9), which is involved in several cancer processes [63,64,65]. This molecule binds near the cleavage site of the MMP-9 zymogen and represents a new pharmacological approach to inhibit this enzyme [66].
In certain cases, however, the goal is just the opposite, namely the enhanced activation of peptidase zymogens. A good example is the family of procaspases, which are cysteine peptidases and key components of apoptosis. Like most peptidases, caspases are expressed in the form of a zymogen. However, their activation process differs from that of the cysteine peptidases mentioned above. There are two subfamilies of caspases involved in the process of cell death: initiator and effector caspases. The initiator caspases, such as procaspase-9 and procaspase-8, are present in the cell in a zymogenic form and undergo autocatalytic cleavage upon receipt of a proapoptotic signal. In addition to autocatalytic cleavage of the zymogen, initiator caspases must form higher complexes with other proteins, such as the apoptosome, to achieve maximal activity [67,68]. Effector caspases, such as procaspase-3, are present in the cell in dimeric zymogen form [69] and are activated by initiator caspases [70]. The initiator caspase cleaves the executor procaspase, allowing the rearrangement of essential loops in the active site. This conformational change leads to the complete formation of the active site and subsequent activation of the peptidase [71]. The activation of effector caspases leads to the proteolytic cleavage of various substrates, resulting in cell death. In the treatment of cancer, inducing the death of cancer cells is desired. Therefore, drugs that stimulate the activation of effector procaspases are being sought. One such molecule is PAC-1, which activates procaspase-3 by forming a tight complex with an inhibitory zinc ion that allows procaspase-3 to activate itself into caspase-3 [72,73]. Another molecule that stimulates procaspase-3 activation is 1541B, which promotes fibril formation and the colocalization of effector and initiator caspases [74,75,76]. These molecules could be used as co-therapies for cancer treatment because they synergistically enhance procaspase-3 activation [77].
Finally, it should be noted that while most peptidases are synthesized as zymogens, there are exceptions to this rule. One such example that is also present in humans is the C1B family, which includes the eukaryotic bleomycin hydrolase [78] that is expressed without a proregion. In addition, viral peptidases are often expressed without a proregion—usually as part of polyprotein chains. This phenomenon can be observed, for example, in herpes simplex virus peptidase 1 (HSV-1) [79], human immunodeficiency virus peptidase 1 (HIV-1) [80], human cytomegalovirus peptidase (hCMV) [81], and the main protease (Mpro) of SARS-CoV-2 [82]. These peptidases become catalytically active after homodimerization or interaction with other modulators. Therefore, a promising strategy for antiviral drugs would be to prevent dimerization of the peptidase (see reviews [83,84]). This would not only solve the problem of antiviral drug resistance but also the problem of lack of specificity of orthosteric drugs. A few studies to prevent dimerization have been performed on HIV-1 and Mpro [85,86,87], but to our knowledge, no such drug is currently in use.

3. Regulation of Peptidase Activity by Glycosaminoglycans

After zymogen activation, peptidase activity can be affected by the presence of macromolecules in its environment, such as glycosaminoglycans (GAGs). GAGs are linear, negatively charged polysaccharides that are ubiquitous components of the extracellular matrix (ECM) as well as internal cellular compartments. They are distinguished from one another by the nature of repeating disaccharide units, degree of sulfation, position of sulfate groups, and resulting negative charge density. In mammals, there are six GAGs, namely chondroitin sulfate (CS), dermatan sulfate (DS), heparin (HP), heparan sulfate (HS), hyaluronan (HA), and keratan sulfate (KS). All are sulfated to some degree and bind to proteins to form proteoglycans, with the exception of HA, which is unsulfated and mostly in a free form [88]. GAGs exert their effects through interactions with various protein targets, including peptidases and their inhibitors.
The most important GAG in the clinic is HP, which has been used for decades as an antithrombotic agent because it accelerates inactivation of the coagulation cascade. The latter involves the sequential activation of serine peptidases, with the final steps of the cascade consisting of the activation of prothrombin, which leads to active thrombin that cleaves fibrinogen to fibrin, the main component of a blood clot. The cascade of the blood-clotting process must be tightly regulated because excessive or insufficient activity of blood clotting factors can lead to thrombosis or hemophilia, respectively. One of the best-studied regulatory mechanisms is the effect of HP on antithrombin, which is a plasma glycoprotein that inhibits the activity of peptidases by forming an equimolar complex with the enzyme [89,90]. When antithrombin binds to HP, it undergoes a conformational change [91] that activates the protein and makes it a better thrombin inhibitor by increasing the accessibility of the inhibitor’s reactive bond to the enzyme (Figure 4). Binding to HP also makes antithrombin a better inhibitor of factor X [92], kallikrein [93,94], plasmin [95], and factor VII [96]. However, the effect of HP on the former factors is less pronounced than its effect on thrombin.
Contrasting effects of GAGs have been demonstrated for serine peptidases stored in the secretory granules of mast cells and neutrophils, including chymase, tryptase, cathepsin G and elastase. The binding of HP to mast cell tryptase is critical for tetramerization and activation of the enzyme [97,98]. When the HP chains are too short to allow tetramerization, they enable the formation of an active monomer [99]. In addition, HP prevents the inhibition of mast cell chymase by its physiological inhibitors (α1-antitrypsin, α2-macroglobulin, and α2-antichymotrypsin) [100] but increases its susceptibility to inhibition by the secretory leukocyte proteinase inhibitor [101]. Similarly, the presence of HP decreases the inhibition of cathepsin G by α1-antichymotrypsin by sterically hindering the binding of the inhibitor [102], but it stimulates its inhibition by the mucus proteinase inhibitor (MPI) [103]. Similar effects have been observed for neutrophil elastase [104,105,106]. Moreover, HP directly inhibits the proteolytic activity of both cathepsin G and neutrophil elastase [107,108,109]. Neutrophil elastase in particular has been studied in detail with regard to its interactions with GAGs. It was found to be inhibited not only by HP but also by other sulfated GAGs. Kinetic experiments have shown that the mechanism of inhibition is typically a hyperbolic mixed inhibition with DS exhibiting the greatest inhibitory effect [110,111,112,113]. Interestingly, GAGs showed a biphasic effect on neutrophil elastase activity; i.e., they acted as inhibitors at low concentrations, but reactivation of the enzyme occurred at higher concentrations, which was attributed to multiple possible binding modes of GAGs to neutrophil elastase [112].
The interactions between GAGs and target proteins are electrostatic, mostly specific and depend on the degree and pattern of sulfation, resulting in different GAGs having different effects on the same target protein [88]. Cathepsin K is a good example of this phenomenon. In particular, chondroitin-4-sulfate (C4S) has the strongest stabilizing effect on this peptidase and increases its collagenolytic activity [114]. The presence of C4S has also been shown to be a necessary prerequisite for the collagenolytic activity of the enzyme, as the main function of the C4S/cathepsin K complex is to partially unfold the triple helix structure of collagen [115,116] by multimeric cathepsin K complexes formed in the presence of C4S [117,118]. Chondroitin-6-sulfate (C6S), which differs from C4S only by the position of the sulfate group at C6 instead of C4, has a 10-fold lower stabilizing effect on cathepsin K [119], demonstrating the specificity of GAGs for their target proteins. GAGs bind to cathepsin K at multiple sites away from the active site and affect enzyme activity by allosteric mechanisms [117,118,120]. The combination of experimental and computational data showed that cathepsin K switches between multiple conformational states and the binding of GAGs stabilizes its more active conformation [120]. In terms of enzyme kinetics, this manifests as the activation (increased activity) of cathepsin K on synthetic substrates [120]. However, differences were observed between the different types of GAGs at neutral pH, as CS and DS acted only as activators, while HP showed signs of diminishing activation effect and even inhibition at high concentrations under the same conditions, reflecting its multiple modes of interaction with cathepsin K [120].
It Is also interesting to note that GAGs have differential effects on closely related papain-like enzymes. While C4S activates cathepsin K, as discussed above, it has an inhibitory effect on the degradation of collagen IV by the closely related lysosomal cathepsin S [121]. Moreover, GAGs can affect the activity of inhibitors of papain-like peptidases but also their inhibitors with contrasting outcomes, similar as discussed above for serine peptidases of immune cells. For example, lysosomal cathepsin L, which is closely related to cathepsins K and S but is not directly regulated by GAGs, is indirectly affected by their interaction with cystatin C, which is a basic low molecular weight protein that acts as a general, emergency inhibitor of papain-like peptidases. In the presence of HP and heparan sulfate (HS), cystatin C cannot bind to the enzyme [122]. In contrast, HP enhances the inhibition of cathepsin L by serpins [123].
As mentioned earlier in this review, GAGs can affect zymogen activation. Studies have demonstrated that GAGs can promote the self-processing of several proenzymes such as procathepsin K [124], procathepsin B [125], and procathepsin S [126]. GAGs also accelerate the activation of certain metalloproteases (e.g., ADAMTS-5, ADAMTS-4, MMP-1, and MMP-2) [127,128,129,130,131] and aspartic peptidases (e.g., cathepsin D) [52].
In a summary of previous research, we have highlighted the complex and diverse effects of GAGs on proteolytic enzymes and emphasized the important role of these molecules in regulating proteolytic activity in both pathological and physiological processes. Therefore, the binding of GAGs to target peptidases represents a potential mechanism for regulating proteolytic activity, but their major limitation is their lack of specificity. Another consideration is that GAGs, as in the case of HP and cathepsin K, can bind to enzymes in different ways, resulting different effects on enzyme activity.

4. Regulation via Exosites

Exosites refer to specific sites of an enzyme that are outside the active site but still play an important role in binding substrates, cofactors, or other ligands. It is important to distinguish exosites from allosteric sites because exosites must normally be occupied for the enzyme to be active. In contrast, binding to the allosteric site only regulates the enzyme’s activity. The significance of exosites in peptidase activity has been studied mainly in the context of the blood coagulation cascade, where all the enzymes involved are trypsin-like serine peptidases. Despite their homology, they exhibit high specificity for biological substrates, which is determined not only by active site residues but also by exosites contributing to substrate affinity. The most important example for analysis of the influence of exosites on proteolytic activity is thrombin, which is the link between the intrinsic and extrinsic pathways of the coagulation cascade. The structure of prothrombin consists of four domains: the Gla domain, two kringle domains involved in protein–protein interactions, and the trypsin-like peptidase domain (Figure 5a). After initiation of the cascade, activated factor Xa cleaves two peptide bonds in prothrombin (Arg320-Ile321 and Arg271-Thr272) and generates active thrombin, which consists of a light and a heavy chain linked by a disulfide bond [132]. Activated thrombin has two distinct exosites that contribute to specific substrate binding and are both positively charged. Exosite I is adjacent to the active site cleft and binds various substrates, cofactors, and inhibitors such as fibrinogen [133], fibrin [134], heparin cofactor II [135], factor XI [136], and thrombomodulin [137]. The second site, referred to as exosite II, which is not an exosite site in the strict sense, as it is more of an allosteric site, is located near the C-terminal helix of the heavy chain. It has been identified as a binding site for HP [138,139] but is now known to bind other regulators such as glycoprotein Ibα [140] (Figure 5b).
Certain natural proteins found in the saliva of bloodsucking and venomous animals also bind to the exosites of thrombin and act as inhibitors or activators of blood coagulation. The first identified peptide of this type is the C-terminal portion of huridin, which is derived from the leech Hirudo meidicinalis and interacts with exosite I (Figure 5c). Although it does not completely inactivate thrombin, the peptide significantly reduces the binding of macromolecular substrates such as fibrinogen, while the peptidase continues to cleave synthetic peptides [141,142]. Bothrojaracin, isolated from the venomous snake Bothops jararaca, achieves a similar effect by binding to both exosites [143,144].
Several other serine peptidases of blood coagulation cascade, such as activated factor X [145], factor IX [146], and factor VII complexed with tissue factor [147], have also identified exosites. The exosite of factor X is also a target of the anticoagulant factor ixolaris, which is derived from the saliva of the tick Ixodes scapularis. Ixolaris, similar to tissue factor pathway inhibitor, binds to the exosite of factor X and inhibits activation by factor VII/tissue factor complex [148].
The presence of exosites has also been demonstrated in cysteine peptidases, including the elastolytic papain-like peptidase cathepsin V, which contains two exosites involved in elastin binding. The mutagenesis of these sites has been shown to result in a loss of elastolytic activity [149], highlighting the importance of exosites in determining enzyme specificity. Similarly, two exosites have been identified in the structure of cathepsin K, both of which are required for effective elastin degradation, whereas only one site is required for the collagenolytic activity of the enzyme. Importantly, the exosites in both cases are not involved in the cleavage of nonstructural substrates [150], making them a promising target for the development of inhibitors that would not interfere with the normal function of the enzyme. An example of such exosite-binding inhibitors is the tanshinones, a group of small plant molecules that selectively inhibit the collagenolytic activity of cathepsin K, demonstrating the potential therapeutic utility of targeting exosites in enzyme inhibition [151,152].
The presence of exosites has also been demonstrated for metalloproteases [153,154,155,156,157,158,159] involved in ECM remodeling and the activation of various biological molecules such as growth factors, cytokines, and factors involved in blood coagulation. Due to their multiple functions, they are involved in the development of various diseases, such as osteoarthritis, thrombosis, bleeding disorders, and cancer (for detailed information on the function of metalloproteases and pathologies, see references [160,161,162,163,164]). Therefore, they represent a promising target for the treatment of these diseases, although specificity remains a major challenge. The development of small molecules and antibodies targeting metalloproteases has therefore focused on binding to exosites, which are the main determinant of the specificity of these proteolytic enzymes.
Researchers have succeeded in producing an antibody that specifically binds to the exosite of ADAMTS5 and inhibits its aggrecanase activity [165], making it a potential treatment option for osteoarthritis. The same peptidase can also be inhibited by glycoconjugated arylsulfonamines [166] and RNA aptamers [167]. Another metalloprotease involved in the development of osteoarthritis is MMP13 [168], whose activity has been successfully inhibited by small molecules with exosite binding [169,170,171,172,173]. Inhibitors in the form of antibodies have also been discovered that selectively inhibit the activity of MMP14, which is involved in the process of cell migration and invasion and whose excessive activity has been linked to the malignancy of various cancers [158,174,175,176]. In addition, specific short peptides have been synthesized that successfully inhibit the activity of MMP2 [177]. These examples suggest that targeting exosites of metalloproteases may be of clinical importance. A more detailed analysis of the inhibition of metalloproteases, including inhibition by exosite binding, can be found in references [178,179,180].
Targeting exosites is also a potential treatment for several infectious diseases, with one of the best-studied examples being Clostridium botulinum infection. The bacterium secretes botulinum neurotoxin serotype A, a zinc-dependent peptidase that specifically cleaves SNAP25 proteins in motor neurons, preventing the release of acetylcholine and leading to paralysis [181]. Currently, there are no drugs that inhibit the activity of the toxin already present in the neuron. In the search for inhibitors, the active site containing a zinc ion must be avoided because this metal is also present in human metalloproteases. For this reason, two specific exosites of the peptidase have shown great potential for inhibition [182,183], and several small molecules [184,185] and antibodies [186,187] have already been discovered that bind to either of the two. Similarly, the lethal metalloprotease of Bacillus anthracis has been inhibited by targeting its exosite [188].

5. Allosteric Regulation by Small Molecule Effectors

Allostery refers to the regulation of enzyme activity involving multiple spatially distant sites that influence each other via conformational changes. Originally, allostery was thought to occur only in oligomeric proteins, and models such as the Monod–Wyman–Changeux model [189] and the Koshland–Nemethy–Filmer model [190] were proposed to describe the relationships between enzyme activity and conformational changes. However, it is now known that allostery is a fundamental property of all dynamic proteins that exist as an ensemble of conformers with different activities and affinities for substrates. Covalent or non-covalent modifications at allosteric sites shift the equilibrium toward a more active or less active conformer. For comprehensive information on allostery in the cell, see the review in ref. [191].
The main advantage of allosteric over orthosteric inhibition (i.e., binding of the inhibitor to the active site in a substrate-like manner) is the ability to fine-tune enzyme activity. Allosteric inhibitors can act as partial inhibitors of peptidase activity by causing a conformational change in the active site that still allows binding of the substrate but with reduced affinity. One of the best-studied human allosteric peptidases is thrombin, which can act as both a procoagulant and an anticoagulant. The procoagulant function of thrombin involves the conversion of fibrinogen to fibrin and an enhancement of its own activity through a positive feedback loop, whereas the anticoagulant function involves the activation of protein C in the presence of thrombomodulin [192]. Thus, the function of thrombin is dependent on its conformation. The enzyme exists in a “fast” and a “slow” conformation, and the transition between these two is dependent on Na+ [193] (Figure 6a). The procoagulant “fast” conformation cleaves fibrinogen with higher specificity and is mainly populated in the presence of Na+ ions, whereas the anticoagulant “slow” conformation, which is more abundant in the absence of Na+ ions, preferentially cleaves protein C [194]. This difference in substrate specificity of the two conformations is due to conformational changes [193,195,196] involving residues in the active site and Na+ binding site [197] (Figure 6b). Thus, the regulation of thrombin activity is complex and is influenced by allosteric communication between the allosteric Na+ binding site, exosites I and II, and the active site.
Allosteric regulation was also observed for other trypsin- and chymotrypsin-like serine peptidases. It was found that proteolytic cleavage of the proregion is not required for zymogen activation. Instead, the binding of an allosteric effector mimicking the activating N-terminal peptide to the allosteric site is sufficient to cause a conformational change from an inactive to an active state [7,198]. The regulation of activation of serine peptidases involved in the blood coagulation cascade is also exploited by pathogens such as streptococci and staphylococci. The former secrete streptokinase, which activates plasminogen allosterically [199], and the latter secrete staphylocoagulase, which activates prothrombin in the same manner [200,201].
Allosteric regulation has also been extensively studied in human papain-like cysteine peptidases, particularly in cathepsin K. The first known allosteric effectors of this enzyme were GAGs, but since then, many other effectors have been discovered that bind to different sites of the protein and have different effects on its activity [202,203]. Interestingly, in the case of cathepsin K, minimal conformational changes occur during allosteric regulation, as shown by molecular dynamics simulations [204]. However, these changes are reflected in modest changes in activity measured with synthetic substrates.
The activity of caspases can be regulated allosterically at multiple stages. The activation of procaspases can be modulated by allosteric inhibitors that bind to the dimerization site and prevent the formation of an active homodimer. In the case of caspase-9, activation is inhibited in this manner by binding of the BIR3 domain of the X-chromosome linked inhibitor of apoptosis (XIAP) to the dimerization site, keeping the peptidase in an inactive conformation [205]. There are also multiple classes of allosteric caspase inhibitors that act at the level of active enzyme, such as FICA, DICA, designed ankyrin repeat proteins (DARPINs) and a class of copper-containing organometallic compounds [206,207,208,209,210]. A common denominator of these inhibitors is their binding in a 2:2 caspase:inhibitor stoichiometry. The small molecule inhibitor DICA and related compounds bind covalently to a conserved Cys residue near the dimer interface (Figure 7a), whereas the organometallic compound NSC321205 binds non-covalently and directly at the dimer interface (Figure 7b).
In addition to the low specificity of orthosteric inhibitors, the advantage of allosteric inhibitors is that they allow fine-tuning of the different enzyme activities of a single target. An example is gamma-secretase, which cleaves multiple substrates, including β-amyloid precursor protein (APP) and Notch1. Cleavage of these substrates has been linked to the development of Alzheimer’s disease (AD) and cancer, respectively, making gamma-secretase an attractive drug target. The inhibition of total enzyme activity proved insufficient for the treatment of AD because toxicity was caused by the simultaneous inhibition of the Notch pathway [211]. Therefore, researchers have focused on developing allosteric effectors that selectively affect substrate affinity [212,213,214,215]. Beta-secretase or BACE1 is also a potential target for the treatment of AD. However, because of the similarity between cathepsin D, gamma-secretase, and BACE1, orthosteric inhibition could lead to off-target toxicity. Therefore, researchers have developed allosteric inhibitors in the form of small molecules, peptides, and antibodies using various computational and experimental methods [216,217,218,219].
In the case of metalloproteases, mainly computational studies have been performed indicating the presence of allosteric sites [220,221,222]. A rare example of demonstrated allostery in metalloproteases is the angiotensin-converting enzyme, which plays a crucial role in blood pressure regulation by hydrolyzing angiotensin I to angiotensin II. In addition, the enzyme has its own domain that cleaves bradykinin. Because these domains have different substrate preferences, allosteric regulation could allow the specific inhibition of only one of the functions, providing a potential avenue for the regulation of blood pressure and fibrosis [223].
Due to increasing bacterial resistance to antibiotics, there is a need for new antibacterial drugs, and bacterial peptidases offer a promising new target. Potential targets include members of the ClpP family, which have been shown to be allosteric (reviewed in reference [224]). These peptidases are responsible for the degradation of defective and misfolded proteins. The ClpP peptidase from Escherichia coli has more than 50 different substrates within the cell and is therefore critical for numerous cellular processes, including the regulation of transcription [225]. Activation of these peptidases would lead to uncontrolled protein degradation, ultimately resulting in cell death of the bacterium. Since members of the same family are also found in human mitochondria, the specificity of allosteric effectors is critical. Allosteric inhibitors may also serve as potential antiviral drugs. For example, the major cysteine peptidase of the SARS-CoV-2 virus [82] has been successfully inhibited allosterically. Most synthetically produced inhibitors bind to the dimerization region and prevent the formation of an active homodimer [226,227,228].

6. Other Examples of Peptidase Regulation Outside of the Active Site

In the previous sections, the examples presented were described primarily from the point of view of the structural basis for their action. However, there are other examples of modifiers that have not been characterized at the structural level and whose kinetic mode of action differs from the specific inhibition associated with orthosteric inhibitors, which is a strong indication that they do not bind to the active site in a substrate-like manner. For example, mixed and catalytic inhibition mechanisms have been diagnosed for the inhibition of cathepsin B by nitroxoline and its derivatives [229,230]. Indeed, nitroxoline has been shown to bind to the S’ sites of the active site of cathepsin B [229]. Similar kinetic mechanisms were discovered for the inhibition of cathepsin B by caffeic acid and chlorogenic acid [231]. However, these two groups responded differently to changes in pH, suggesting that their mode of action is not identical [231].
With complete inhibition, peptidase action is completely suppressed, which is not optimal for cellular functions. However, with partial inhibition, we can maintain the basic activity required for homeostasis. Partial (hyperbolic) inhibition mechanisms were described for allosteric effectors of cathepsin K NSC1334 and NSC94914 [202,203] as well as putative such modifiers of cathepsin K and cathepsin S [203,232]. Several of them, including NSC13345 and NSC94914, showed peculiar kinetic profiles of inhibition and/or activation, depending on the concentration of the substrate [202,203], and altogether, a diverse spectrum of effects ranging from partial specific to partial catalytic inhibition have been observed [203]. Similarly, sulfated coumarins, allosteric inhibitors of thrombin, have been shown to act as partial inhibitors of this peptidase [233]. Partial inhibition is not limited to small molecule modifiers. Various metal ions can partially inhibit not only the activity of cathepsin K but also that of cathepsin S [234]. At the other end of the spectrum, it can also be obtained with proteinaceous binding partners of peptidases that bind away from the active site. Good examples of such action are DARPINs that bind to and inhibit caspase-2 [207] and cathepsin B [235], respectively. The former case is particularly interesting as only 4% of catalytic activity of the peptidase are retained in the complex, and careful kinetic analysis was necessary to discriminate its mode of action from linear inhibition [207].

7. Non-Inhibitory Binding Partners of Peptidases

A wholly distinct type of regulation of peptidase activity, which is also noteworthy in the context of this review, is the interactions of peptidases with non-inhibitory binding partners. These are typically proteins that bind the peptidase outside its active site without having any appreciable direct effect on the kinetic parameters of enzyme activity. Nevertheless, their presence may influence other aspects of the enzymatic activity of the peptidase, such as stability, localization, etc. One of the best-known categories of such binding partners is membrane proteins, which act as specific receptors for secreted peptidases. A large repertoire of such receptors is available for MMPs, including integrins, tetraspanins, emmprin, etc. (reviewed by Yamamoto et al. [236]). These interactions have been identified for both soluble MMPs [237,238] and membrane-type MMPs, most notably MT-MMP1 [239]. They have been shown to regulate both pericellular proteolysis [237,239] as well as the trafficking and recycling of these peptidases [240]. Well-characterized receptors are also known for the serine peptidase plasmin(ogen) and its activator urokinase. Together, the plasminogen and urokinase plasminogen activator receptors (uPARs) contribute to the activation of plasmin at the cell surface and the concomitant proteolytic destruction of the pericellular extracellular matrix that is the hallmark of many cancers [241]. Not surprisingly, these interactions are considered potential targets for anticancer therapies [242] as well as diagnostic imaging [243]. Similarly, cysteine cathepsin B, an otherwise unstable peptidase in the extracellular environment [244] which is also associated with cancer progression [245], was found to mediate pericellular proteolysis by binding to caveolin-1 [246,247] and annexin II [248]. Unfortunately, this type of interaction has remained uninvestigated for most peptidases and thus potentially representing a significant untapped potential for drug targeting that is worth exploring in the future.
Alternatively, interactions with non-inhibitory binding partners may mediate non-enzymatic physiological and/or pathological functions of peptidases. A well-known example is the transmembrane form of dipeptidyl-peptidase 4 (DPP4), an important target in type 2 diabetes, which not only regulates immune cells via its enzymatic activity but also functions as a cell surface receptor (reviewed in [249]). For example, its interactions with extracellular adenosine deaminase [250] and CD45 [251] regulate the activation of T cells. In addition, DPP4 has been identified as the entry receptor for Middle East respiratory syndrome (MERS) coronavirus (CoV) [252]. Another much more infamous example of peptidases as viral entry points is the cell surface metalloprotease angiotensin-converting enzyme 2 (ACE2), which serves as the receptor used by SARS-CoV [253,254]. Disruption of this interaction by neutralizing antibodies has been identified as one of the key elements for a successful immune response to infection [255]. It is therefore not surprising that ACE2 mimetics, which act as scavengers of viral particles, are emerging as treatment for SARS-CoV infections [256]. Non-proteolytic functions of peptidases may also contribute to the pathogenicity of various microbial pathogens. In this case, the culprits are microbial peptidases that contribute to host invasion (reviewed by Jarocki et al. [257]). Examples include streptopain (SpeB) and C5a from streptococci, which mediate the adhesion of bacterial cells to structural proteins of the extracellular matrix [258,259], plasminogen activator from Yersinia pestis, etc. Such interactions, too, are potentially druggable and may provide novel opportunities for disease management.

8. Conclusions

Many peptidases have been implicated in various human diseases characterized by either up- or down-regulation of their catalytic activity. Given their involvement in pathological processes, the study of the regulation of proteolytic activity has attracted considerable interest. Moreover, proteolytic enzymes of pathogens are potential targets for new generations of antibiotics. Looking at the development of drugs in clinical trials, it is reasonable to assume that most of the drugs currently being tested will unfortunately not reach clinical application [260]. The problem usually lies in the specificity of the drugs, which not only bind to the target protein but also have unexpected interactions with other peptidases [261]. To overcome this challenge, recent research has focused on finding novel ways to regulate peptidase activity, such as by regulating zymogen activation, modulating enzyme activity via the binding of effectors to exosites, and allosteric sites. These mechanisms offer potential opportunities for targeting human, viral, and bacterial proteolytic enzymes. However, it is critical to note that the regulation of peptidase activity is a complex process that results from the sum of multiple mechanisms, as seen in the case of thrombin.

Author Contributions

Writing—original draft preparation, A.O.; writing—review and editing, M.N. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Slovenian Research and Innovation Agency, grant number N1-0211 and research core funding number P1-0179.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Rawlings, N.D.; Barrett, A.J.; Thomas, P.D.; Huang, X.; Bateman, A.; Finn, R.D. The MEROPS Database of Proteolytic Enzymes, Their Substrates and Inhibitors in 2017 and a Comparison with Peptidases in the PANTHER Database. Nucleic Acids Res. 2018, 46, D624–D632. [Google Scholar] [CrossRef] [PubMed]
  2. Zhou, Y.; Zhang, Y.; Lian, X.; Li, F.; Wang, C.; Zhu, F.; Qiu, Y.; Chen, Y. Therapeutic Target Database Update 2022: Facilitating Drug Discovery with Enriched Comparative Data of Targeted Agents. Nucleic Acids Res. 2022, 50, D1398–D1407. [Google Scholar] [CrossRef]
  3. Tanaka, K.A.; Levy, J.H. Regulation of Thrombin Activity—Pharmacologic and Structural Aspects. Hematol. Oncol. Clin. N. Am. 2007, 21, 33–50. [Google Scholar] [CrossRef] [PubMed]
  4. Scheen, A.J. The Safety of Gliptins : Updated Data in 2018. Expert Opin. Drug Saf. 2018, 17, 387–405. [Google Scholar] [CrossRef]
  5. McClung, M.R.; O’Donoghue, M.L.; Papapoulos, S.E.; Bone, H.; Langdahl, B.; Saag, K.G.; Reid, I.R.; Kiel, D.P.; Cavallari, I.; Bonaca, M.P.; et al. Odanacatib for the Treatment of Postmenopausal Osteoporosis: Results of the LOFT Multicentre, Randomised, Double-Blind, Placebo-Controlled Trial and LOFT Extension Study. Lancet Diabetes Endocrinol. 2019, 7, 899–911. [Google Scholar] [CrossRef] [PubMed]
  6. Freer, S.T.; Kraut, J.; Robertus, J.D.; Wright, H.T.; Xuong, N.H. Chymotrypsinogen: 2.5-Angstrom Crystal Structure, Comparison with Alpha-Chymotrypsin, and Implications for Zymogen Activation. Biochemistry 1970, 9, 1997–2009. [Google Scholar] [CrossRef]
  7. Bode, W.; Schwager, P.; Huber, R. The Transition of Bovine Trypsinogen to a Trypsin-like State upon Strong Ligand Binding. The Refined Crystal Structures of the Bovine Trypsinogen-Pancreatic Trypsin Inhibitor Complex and of Its Ternary Complex with Ile-Val at 1.9 A Resolution. J. Mol. Biol. 1978, 118, 99–112. [Google Scholar] [CrossRef]
  8. Birktoft, J.J.; Kraut, J.; Freer, S.T. A Detailed Structural Comparison between the Charge Relay System in Chymotrypsinogen and in Alpha-Chymotrypsin. Biochemistry 1976, 15, 4481–4485. [Google Scholar] [CrossRef]
  9. Daviet, E.W.; Neurath, H. Identification of a peptide released during autocatalytic activation of trypsinogen. J. Biol. Chem. 1955, 212, 515–529. [Google Scholar] [CrossRef]
  10. Russo, S.F.; Wahl, E.T. Autocatalytic Activation of Trypsinogen by Trypsin. J. Chem. Educ. 1987, 64, 83–84. [Google Scholar] [CrossRef]
  11. Yamashina, I. The Action of Enterokinase on Trypsinogen. Biochim. Biophys. Acta 1956, 20, 433–434. [Google Scholar] [CrossRef]
  12. Drbybr, W.J.; Neurath, H. The Activation of Chymotrypsinogen. J. Am. Chem. Soc. 1955, 77, 814–815. [Google Scholar] [CrossRef]
  13. Pham, C.T.N.; Ley, T.J. Dipeptidyl Peptidase I Is Required for the Processing and Activation of Granzymes A and B in Vivo. Proc. Natl. Acad. Sci. USA 1999, 96, 8627–8632. [Google Scholar] [CrossRef] [PubMed]
  14. Adkison, A.M.; Raptis, S.Z.; Kelley, D.G.; Pham, C.T.N. Dipeptidyl Peptidase I Activates Neutrophil-Derived Serine Proteases and Regulates the Development of Acute Experimental Arthritis. J. Clin. Investig. 2002, 109, 363–371. [Google Scholar] [CrossRef] [PubMed]
  15. Wolters, P.J.; Pham, C.T.N.; Muilenburg, D.J.; Ley, T.J.; Caughey, G.H. Dipeptidyl Peptidase I Is Essential for Activation of Mast Cell Chymases, but Not Tryptases, in Mice. J. Biol. Chem. 2001, 276, 18551–18556. [Google Scholar] [CrossRef]
  16. Stojkovska Docevska, M.; Novinec, M. Cathepsin C: Structure, Function, and Pharmacological Targeting. Rare Dis. Orphan Drugs J. 2023, 2, 14. [Google Scholar] [CrossRef]
  17. Cipolla, D.; Zhang, J.; Korkmaz, B.; Chalmers, J.D.; Basso, J.; Lasala, D.; Fernandez, C.; Teper, A.; Mange, K.C.; Perkins, W.R.; et al. Dipeptidyl Peptidase-1 Inhibition with Brensocatib Reduces the Activity of All Major Neutrophil Serine Proteases in Patients with Bronchiectasis: Results from the WILLOW Trial. Respir. Res. 2023, 24, 133. [Google Scholar] [CrossRef] [PubMed]
  18. Yanamandra, N.; Gumidyala, K.V.; Waldron, K.G.; Gujrati, M.; Olivero, W.C.; Dinh, D.H.; Rao, J.S.; Mohanam, S. Blockade of Cathepsin B Expression in Human Glioblastoma Cells Is Associated with Suppression of Angiogenesis. Oncogene 2004, 23, 2224–2230. [Google Scholar] [CrossRef]
  19. Moon, H.Y.; Becke, A.; Berron, D.; Becker, B.; Sah, N.; Benoni, G.; Janke, E.; Lubejko, S.T.; Greig, N.H.; Mattison, J.A.; et al. Running-Induced Systemic Cathepsin B Secretion Is Associated with Memory Function. Cell Metab. 2016, 24, 332–340. [Google Scholar] [CrossRef]
  20. Lakka, S.S.; Gondi, C.S.; Yanamandra, N.; Olivero, W.C.; Dinh, D.H.; Gujrati, M.; Rao, J.S. Inhibition of Cathepsin B and MMP-9 Gene Expression in Glioblastoma Cell Line via RNA Interference Reduces Tumor Cell Invasion, Tumor Growth and Angiogenesis. Oncogene 2004, 23, 4681–4689. [Google Scholar] [CrossRef]
  21. Veeravalli, K.K.; Chetty, C.; Ponnala, S.; Gondi, C.S.; Lakka, S.S.; Fassett, D.; Klopfenstein, J.D.; Dinh, D.H.; Gujrati, M.; Rao, J.S. MMP-9, UPAR and Cathepsin B Silencing Downregulate Integrins in Human Glioma Xenograft Cells in Vitro and in Vivo in Nude Mice. PLoS ONE 2010, 5, e11583. [Google Scholar] [CrossRef]
  22. Quraishi, O.; Nägler, D.K.; Fox, T.; Sivaraman, J.; Cygler, M.; Mort, J.S.; Storer, A.C. The Occluding Loop in Cathepsin B Defines the PH Dependence of Inhibition by Its Propeptide. Biochemistry 1999, 38, 5017–5023. [Google Scholar] [CrossRef]
  23. Pungercar, J.R.; Caglic, D.; Sajid, M.; Dolinar, M.; Vasiljeva, O.; Pozgan, U.; Turk, D.; Bogyo, M.; Turk, V.; Turk, B. Autocatalytic Processing of Procathepsin B Is Triggered by Proenzyme Activity. FEBS J. 2009, 276, 660–668. [Google Scholar] [CrossRef]
  24. Rozman, J.; Stojan, J.; Kuhelj, R.; Turk, V.; Turk, B. Autocatalytic Processing of Recombinant Human Procathepsin B Is a Bimolecular Process. FEBS Lett. 1999, 459, 358–362. [Google Scholar] [CrossRef]
  25. McQueney, M.S.; Amegadzie, B.Y.; D’Alessio, K.; Hanning, C.R.; McLaughlin, M.M.; McNulty, D.; Carr, S.A.; Ijames, C.; Kurdyla, J.; Jones, C.S. Autocatalytic Activation of Human Cathepsin K. J. Biol. Chem. 1997, 272, 13955–13960. [Google Scholar] [CrossRef] [PubMed]
  26. Ishidoh, K.; Kominami, E. Multi-Step Processing of Procathepsin L in Vitro. FEBS Lett. 1994, 352, 281–284. [Google Scholar] [CrossRef] [PubMed]
  27. Nomura, T.; Fujisawa, Y. Processing Properties of Recombinant Human Procathepsin L. Biochem. Biophys. Res. Commun. 1997, 230, 143–146. [Google Scholar] [CrossRef]
  28. Ménard, R.; Carmona, E.; Takebe, S.; Dufour, É.; Plouffe, C.; Mason, P.; Mort, J.S. Autocatalytic Processing of Recombinant Human Procathepsin L. Contribution of Both Intermolecular and Unimolecular Events in the Processing of Procathepsin L in Vitro. J. Biol. Chem. 1998, 273, 4478–4484. [Google Scholar] [CrossRef]
  29. Kopitar, G.; Dolinar, M.; Štrukelj, B.; Pungerčar, J.; Turk, V. Folding and Activation of Human Procathepsin S from Inclusion Bodies Produced in Escherichia Coli. Eur. J. Biochem. 1996, 236, 558–562. [Google Scholar] [CrossRef] [PubMed]
  30. Wang, B.; Shi, G.P.; Yao, P.M.; Li, Z.; Chapman, H.A.; Brömme, D. Human Cathepsin F. Molecular Cloning, Functional Expression, Tissue Localization, and Enzymatic Characterization. J. Biol. Chem. 1998, 273, 32000–32008. [Google Scholar] [CrossRef]
  31. Sivaraman, J.; Nägler, D.K.; Zhang, R.; Ménard, R.; Cygler, M. Crystal Structure of Human Procathepsin X: A Cysteine Protease with the Proregion Covalently Linked to the Active Site Cysteine. J. Mol. Biol. 2000, 295, 939–951. [Google Scholar] [CrossRef] [PubMed]
  32. Nägler, D.K.; Zhang, R.; Tam, W.; Sulea, T.; Purisima, E.O.; Ménard, R. Human Cathepsin X: A Cysteine Protease with Unique Carboxypeptidase Activity. Biochemistry 1999, 38, 12648–12654. [Google Scholar] [CrossRef] [PubMed]
  33. Turk, D.; Janjic, V.; Stern, I.; Podobnik, M.; Lamba, D.; Dahl, S.W.; Lauritzen, C.; Pedersen, J.; Turk, V.; Turk, B. Structure of Human Dipeptidyl Peptidase I (Cathepsin C): Exclusion Domain Added to an Endopeptidase Framework Creates the Machine for Activation of Granular Serine Proteases. EMBO J. 2001, 20, 6570–6582. [Google Scholar] [CrossRef]
  34. Rebernik, M.; Lenarčič, B.; Novinec, M. The Catalytic Domain of Cathepsin C (Dipeptidyl-Peptidase I) Alone Is a Fully Functional Endoprotease. Protein Expr. Purif. 2019, 157, 21–27. [Google Scholar] [CrossRef] [PubMed]
  35. Dahl, S.W.; Halkier, T.; Lauritzen, C.; Dolenc, I.; Pedersen, J.; Turk, V.; Turk, B. Human Recombinant Pro-Dipeptidyl Peptidase I (Cathepsin C) Can Be Activated by Cathepsins L and S but Not by Autocatalytic Processing. Biochemistry 2001, 40, 1671–1678. [Google Scholar] [CrossRef]
  36. Gunčar, G.; Podobnik, M.; Pungerčar, J.; Štrukelj, B.; Turk, V.; Turk, D. Crystal Structure of Porcine Cathepsin H Determined at 2.1 å Resolution: Location of the Mini-Chain C-Terminal Carboxyl Group Defines Cathepsin H Aminopeptidase Function. Structure 1998, 6, 51–61. [Google Scholar] [CrossRef] [PubMed]
  37. Hao, Y.; Purtha, W.; Cortesio, C.; Rui, H.; Gu, Y.; Chen, H.; Sickmier, E.A.; Manzanillo, P.; Huang, X. Crystal Structures of Human Procathepsin H. PLoS ONE 2018, 13, e0200374. [Google Scholar] [CrossRef]
  38. Van Wart, H.E.; Birkedal-Hansen, H. The Cysteine Switch: A Principle of Regulation of Metalloproteinase Activity with Potential Applicability to the Entire Matrix Metalloproteinase Gene Family. Proc. Natl. Acad. Sci. USA 1990, 87, 5578–5582. [Google Scholar] [CrossRef]
  39. Chen, L.C.; Noelken, M.E.; Nagase, H. Disruption of the Cysteine-75 and Zinc Ion Coordination Is Not Sufficient to Activate the Precursor of Human Matrix Metalloproteinase 3 (Stromelysin 1). Biochemistry 1993, 32, 10289–10295. [Google Scholar] [CrossRef]
  40. Nagase, H.; Enghild, J.J.; Suzuki, K.; Salvesen, G. Stepwise Activation Mechanisms of the Precursor of Matrix Metalloproteinase 3 (Stromelysin) by Proteinases and (4-Aminophenyl)Mercuric Acetate. Biochemistry 1990, 29, 5783–5789. [Google Scholar] [CrossRef]
  41. Peppin, G.J.; Weiss, S.J. Activation of the Endogenous Metalloproteinase, Gelatinase, by Triggered Human Neutrophils. Proc. Natl. Acad. Sci. USA 1986, 83, 4322–4326. [Google Scholar] [CrossRef] [PubMed]
  42. Weiss, S.J.; Peppin, G.; Ortiz, X.; Ragsdale, C.; Test, S.T. Oxidative Autoactivation of Latent Collagenase by Human Neutrophils. Science 1985, 227, 747–749. [Google Scholar] [CrossRef] [PubMed]
  43. Fu, X.; Kassimm, S.Y.; Parks, W.C.; Heinecke, J.W. Hypochlorous Acid Oxygenates the Cysteine Switch Domain of Pro-Matrilysin (MMP-7). A Mechanism for Matrix Metalloproteinase Activation and Atherosclerotic Plaque Rupture by Myeloperoxidase. J. Biol. Chem. 2001, 276, 41279–41287. [Google Scholar] [CrossRef] [PubMed]
  44. Perlmann, G.E. The Optical Rotatory Properties of Pepsinogen. J. Mol. Biol. 1963, 6, 452–464. [Google Scholar] [CrossRef] [PubMed]
  45. James, M.N.G.; Sielecki, A.R. Molecular Structure of an Aspartic Proteinase Zymogen, Porcine Pepsinogen, at 1.8 A Resolution. Nature 1986, 319, 33–38. [Google Scholar] [CrossRef] [PubMed]
  46. Al-Janabi, J.; Hartsuck, J.A.; Tang, J. Kinetics and Mechanism of Pepsinogen Activation. J. Biol. Chem. 1972, 247, 4628–4632. [Google Scholar] [CrossRef]
  47. Derkx, F.H.; Schalekamp, M.P.; Schalekamp, M.A. Two-Step Prorenin-Renin Conversion. Isolation of an Intermediary Form of Activated Prorenin. J. Biol. Chem. 1987, 262, 2472–2477. [Google Scholar] [CrossRef]
  48. Gieselmann, V.; Hasilik, A.; Von Figura, K. Processing of Human Cathepsin D in Lysosomes in Vitro. J. Biol. Chem. 1985, 260, 3215–3220. [Google Scholar] [CrossRef]
  49. Hasilik, A.; von Figura, K.; Conzelmann, E.; Nehrkorn, H.; Sandhoff, K. Lysosomal Enzyme Precursors in Human Fibroblasts. Activation of Cathepsin D Precursor in Vitro and Activity of Beta-Hexosaminidase A Precursor towards Ganglioside GM2. Eur. J. Biochem. 1982, 125, 317–321. [Google Scholar] [CrossRef]
  50. Richo, G.R.; Conner, G.E. Structural Requirements of Procathepsin D Activation and Maturation. J. Biol. Chem. 1994, 269, 14806–14812. [Google Scholar] [CrossRef]
  51. Wittlin, S.; Ro, È.; Sel, J.; Hofmann, F.; Stover, D.R. Mechanisms and Kinetics of Procathepsin D Activation. Eur. J. Biochem. 1999, 265, 384–393. [Google Scholar] [CrossRef]
  52. Beckman, M.; Freeman, C.; Parish, C.R.; Small, D.H. Activation of Cathepsin D by Glycosaminoglycans. FEBS J. 2009, 276, 7343–7352. [Google Scholar] [CrossRef]
  53. Fox, T.; Storer, A.C.; de Miguel, E.; Mort, J.S. Potent Slow-Binding Inhibition of Cathepsin B by Its Propeptide. Biochemistry 1992, 31, 12571–12576. [Google Scholar] [CrossRef]
  54. Carmona, E.; Dufour, É.; Plouffe, C.; Takebe, S.; Mason, P.; Mort, J.S.; Ménard, R. Potency and Selectivity of the Cathepsin L Propeptide as an Inhibitor of Cysteine Proteases. Biochemistry 1996, 35, 8149–8157. [Google Scholar] [CrossRef] [PubMed]
  55. Maubach, G.; Schilling, K.; Rommerskirch, W.; Wenz, I.; Schultz, J.E.; Weber, E.; Wiederanders, B. The Inhibition of Cathepsin S by Its Propeptide--Specificity and Mechanism of Action. Eur. J. Biochem. 1997, 250, 745–750. [Google Scholar] [CrossRef] [PubMed]
  56. Guay, J.; Falgueyret, J.P.; Ducret, A.; Percival, M.D.; Mancini, J.A. Potency and Selectivity of Inhibition of Cathepsin K, L and S by Their Respective Propeptides. Eur. J. Biochem. 2000, 267, 6311–6318. [Google Scholar] [CrossRef] [PubMed]
  57. Reichhardt, M.P.; Johnson, S.; Tang, T.; Morgan, T.; Tebeka, N.; Popitsch, N.; Deme, J.C.; Jore, M.M.; Lea, S.M. An Inhibitor of Complement C5 Provides Structural Insights into Activation. Proc. Natl. Acad. Sci. USA 2020, 117, 362–370. [Google Scholar] [CrossRef] [PubMed]
  58. Xu, Z.; Chen, Y.W.; Battu, A.; Wilder, P.; Weber, D.; Yu, W.; MacKerell, A.D.; Chen, L.M.; Chai, K.X.; Johnson, M.D.; et al. Targeting Zymogen Activation to Control the Matriptase-Prostasin Proteolytic Cascade. J. Med. Chem. 2011, 54, 7567–7578. [Google Scholar] [CrossRef]
  59. Lee, S.L.; Dickson, R.B.; Lin, C.Y. Activation of Hepatocyte Growth Factor and Urokinase/Plasminogen Activator by Matriptase, an Epithelial Membrane Serine Protease. J. Biol. Chem. 2000, 275, 36720–36725. [Google Scholar] [CrossRef]
  60. List, K.; Szabo, R.; Molinolo, A.; Sriuranpong, V.; Redeye, V.; Murdock, T.; Burke, B.; Nielsen, B.S.; Gutkind, J.S.; Bugge, T.H. Deregulated Matriptase Causes Ras-Independent Multistage Carcinogenesis and Promotes Ras-Mediated Malignant Transformation. Genes Dev. 2005, 19, 1934–1950. [Google Scholar] [CrossRef]
  61. Andreasen, P.A.; Egelund, R.; Petersen, H.H. The Plasminogen Activation System in Tumor Growth, Invasion, and Metastasis. Cell. Mol. Life Sci. 2000, 57, 25–40. [Google Scholar] [CrossRef] [PubMed]
  62. Blouse, G.E.; Bøotkjaer, K.A.; Deryugina, E.; Byszuk, A.A.; Jensen, J.M.; Mortensen, K.K.; Quigley, J.P.; Andreasen, P.A. A Novel Mode of Intervention with Serine Protease Activity: Targeting Zymogen Activation. J. Biol. Chem. 2009, 284, 4647–4657. [Google Scholar] [CrossRef]
  63. Bergers, G.; Brekken, R.; McMahon, G.; Vu, T.H.; Itoh, T.; Tamaki, K.; Tanzawa, K.; Thorpe, P.; Itohara, S.; Werb, Z.; et al. Matrix Metalloproteinase-9 Triggers the Angiogenic Switch during Carcinogenesis. Nat. Cell Biol. 2000, 2, 737–744. [Google Scholar] [CrossRef]
  64. Song, Z.; Wang, J.; Su, Q.; Luan, M.; Chen, X.; Xu, X. The Role of MMP-2 and MMP-9 in the Metastasis and Development of Hypopharyngeal Carcinoma. Braz. J. Otorhinolaryngol. 2021, 87, 521–528. [Google Scholar] [CrossRef] [PubMed]
  65. Mehner, C.; Hockla, A.; Miller, E.; Ran, S.; Radisky, D.C.; Radisky, E.S. Tumor Cell-Produced Matrix Metalloproteinase 9 (MMP-9) Drives Malignant Progression and Metastasis of Basal-like Triple Negative Breast Cancer. Oncotarget 2014, 5, 2736–2749. [Google Scholar] [CrossRef] [PubMed]
  66. Scannevin, R.H.; Alexander, R.; Haarlander, T.M.; Burke, S.L.; Singer, M.; Huo, C.; Zhang, Y.M.; Maguire, D.; Spurlino, J.; Deckman, I.; et al. Discovery of a Highly Selective Chemical Inhibitor of Matrix Metalloproteinase-9 (MMP-9) That Allosterically Inhibits Zymogen Activation. J. Biol. Chem. 2017, 292, 17963–17974. [Google Scholar] [CrossRef]
  67. Muzio, M.; Stockwell, B.R.; Stennicke, H.R.; Salvesen, G.S.; Dixit, V.M. An Induced Proximity Model for Caspase-8 Activation. J. Biol. Chem. 1998, 273, 2926–2930. [Google Scholar] [CrossRef]
  68. Li, P.; Nijhawan, D.; Budihardjo, I.; Srinivasula, S.M.; Ahmad, M.; Alnemri, E.S.; Wang, X. Cytochrome c and DATP-Dependent Formation of Apaf-1/Caspase-9 Complex Initiates an Apoptotic Protease Cascade. Cell 1997, 91, 479–489. [Google Scholar] [CrossRef]
  69. Pop, C.; Chen, Y.R.; Smith, B.; Bose, K.; Bobay, B.; Tripathy, A.; Franzen, S.; Clark, A.C. Removal of the Pro-Domain Does Not Affect the Conformation of the Procaspase-3 Dimer. Biochemistry 2001, 40, 14224–14235. [Google Scholar] [CrossRef]
  70. Stennicke, H.R.; Jürgensmeier, J.M.; Shin, H.; Deveraux, Q.; Wolf, B.B.; Yang, X.; Zhou, Q.; Ellerby, H.M.; Ellerby, L.M.; Bredesen, D.; et al. Pro-Caspase-3 Is a Major Physiologic Target of Caspase-8. J. Biol. Chem. 1998, 273, 27084–27090. [Google Scholar] [CrossRef]
  71. Chai, J.; Wu, Q.; Shiozaki, E.; Srinivasula, S.M.; Alnemri, E.S.; Shi, Y. Crystal Structure of a Procaspase-7 Zymogen: Mechanisms of Activation and Substrate Binding. Cell 2001, 107, 399–407. [Google Scholar] [CrossRef]
  72. Peterson, Q.P.; Goode, D.R.; West, D.C.; Ramsey, K.N.; Lee, J.J.Y.; Hergenrother, P.J. PAC-1 Activates Procaspase-3 in Vitro through Relief of Zinc-Mediated Inhibition. J. Mol. Biol. 2009, 388, 144. [Google Scholar] [CrossRef] [PubMed]
  73. Putt, K.S.; Chen, G.W.; Pearson, J.M.; Sandhorst, J.S.; Hoagland, M.S.; Kwon, J.T.; Hwang, S.K.; Jin, H.; Churchwell, M.I.; Cho, M.H.; et al. Small-Molecule Activation of Procaspase-3 to Caspase-3 as a Personalized Anticancer Strategy. Nat. Chem. Biol. 2006, 2, 543–550. [Google Scholar] [CrossRef] [PubMed]
  74. Wolan, D.W.; Zorn, J.A.; Gray, D.C.; Wells, J.A. Small-Molecule Activators of a Proenzyme. Science 2009, 326, 853–858. [Google Scholar] [CrossRef] [PubMed]
  75. Zorn, J.A.; Wille, H.; Wolan, D.W.; Wells, J.A. Self-Assembling Small Molecules Form Nanofibrils That Bind Procaspase-3 to Promote Activation. J. Am. Chem. Soc. 2011, 133, 19630–19633. [Google Scholar] [CrossRef] [PubMed]
  76. Zorn, J.A.; Wolan, D.W.; Agard, N.J.; Wells, J.A. Fibrils Colocalize Caspase-3 with Procaspase-3 to Foster Maturation. J. Biol. Chem. 2012, 287, 33781–33795. [Google Scholar] [CrossRef] [PubMed]
  77. Botham, R.C.; Fan, T.M.; Im, I.; Borst, L.B.; Dirikolu, L.; Hergenrother, P.J. Dual Small-Molecule Targeting of Procaspase-3 Dramatically Enhances Zymogen Activation and Anticancer Activity. J. Am. Chem. Soc. 2014, 136, 1312–1319. [Google Scholar] [CrossRef]
  78. O’Farrell, P.A.; Gonzalez, F.; Zheng, W.; Johnston, S.A.; Joshua-Tor, L. Crystal Structure of Human Bleomycin Hydrolase, a Self-Compartmentalizing Cysteine Protease. Structure 1999, 7, 619–627. [Google Scholar] [CrossRef]
  79. Schmidt, U.; Darke, P.L. Dimerization and Activation of the Herpes Simplex Virus Type 1 Protease. J. Biol. Chem. 1997, 272, 7732–7735. [Google Scholar] [CrossRef]
  80. Wlodawer, A.; Miller, M.; Jaskólski, M.; Sathyanarayana, B.K.; Baldwin, E.; Weber, I.T.; Selk, L.M.; Clawson, L.; Schneider, J.; Kent, S.B.H. Conserved Folding in Retroviral Proteases: Crystal Structure of a Synthetic HIV-1 Protease. Science 1989, 245, 616–621. [Google Scholar] [CrossRef]
  81. Darke, P.L.; Cole, J.L.; Waxman, L.; Hall, D.L.; Sardana, M.K.; Kuo, L.C. Active Human Cytomegalovirus Protease Is a Dimer. J. Biol. Chem. 1996, 271, 7445–7449. [Google Scholar] [CrossRef]
  82. Zhang, L.; Lin, D.; Sun, X.; Curth, U.; Drosten, C.; Sauerhering, L.; Becker, S.; Rox, K.; Hilgenfeld, R. Crystal Structure of SARS-CoV-2 Main Protease Provides a Basis for Design of Improved α-Ketoamide Inhibitors. Science 2020, 368, 409. [Google Scholar] [CrossRef] [PubMed]
  83. Majerová, Á.; Novotný, P. Precursors of Viral Proteases as Distinct Drug Targets. Viruses 2021, 13, 1981. [Google Scholar] [CrossRef] [PubMed]
  84. Zephyr, J.; Kurt Yilmaz, N.; Schiffer, C.A. Viral Proteases: Structure, Mechanism and Inhibition. Enzymes 2021, 50, 301. [Google Scholar] [CrossRef]
  85. Bannwarth, L.; Rose, T.; Dufau, L.; Vanderesse, R.; Dumond, J.; Jamart-Grégoire, B.; Pannecouque, C.; De Clercq, E.; Le Reboud-Ravaux, M. Dimer Disruption and Monomer Sequestration by Alkyl Tripeptides Are Successful Strategies for Inhibiting Wild-Type and Multidrug-Resistant Mutated HIV-1 Proteases. Biochemistry 2008, 48, 379–387. [Google Scholar] [CrossRef] [PubMed]
  86. Zutshi, R.; Franciskovich, J.; Shultz, M.; Schweitzer, B.; Bishop, P.; Wilson, M.; Chmielewski, J. Targeting the Dimerization Interface of HIV-1 Protease: Inhibition with Cross-Linked Interfacial Peptides. J. Am. Chem. Soc. 1997, 119, 4841–4845. [Google Scholar] [CrossRef]
  87. Goyal, B.; Goyal, D. Targeting the Dimerization of the Main Protease of Coronaviruses: A Potential Broad-Spectrum Therapeutic Strategy. ACS Comb. Sci. 2020, 22, 297–305. [Google Scholar] [CrossRef]
  88. Vallet, S.D.; Clerc, O.; Ricard-Blum, S. Glycosaminoglycan-Protein Interactions: The First Draft of the Glycosaminoglycan Interactome. J. Histochem. Cytochem. 2021, 69, 93–104. [Google Scholar] [CrossRef]
  89. Abildgaard, U. Binding of Thrombin to Antithrombin III. Scand. J. Clin. Lab. Investig. 1969, 24, 23–27. [Google Scholar] [CrossRef]
  90. Rosenberg, R.D.; Damus, P.S. The Purification and Mechanism of Action of Human Antithrombin-Heparin Cofactor. J. Biol. Chem. 1973, 248, 6490–6505. [Google Scholar] [CrossRef]
  91. Jin, L.; Abrahams, J.P.; Skinner, R.; Petitou, M.; Pike, R.N.; Carrell, R.W. The Anticoagulant Activation of Antithrombin by Heparin. Proc. Natl. Acad. Sci. USA 1997, 94, 14683–14688. [Google Scholar] [CrossRef] [PubMed]
  92. Craig, P.A.; Olson, S.T.; Shore, J.D. Transient Kinetics of Heparin-Catalyzed Protease Inactivation by Antithrombin III: Characterization of Assembly, Product Formation, and Heparin Dissociation Steps in the Factor Xa Reaction. J. Biol. Chem. 1989, 264, 5452–5461. [Google Scholar] [CrossRef] [PubMed]
  93. Gozzo, A.J.; Nunes, V.A.; Nader, H.B.; Dietrich, C.P.; Carmona, A.K.; Sampaio, M.U.; Sampaio, C.A.M.; Araújo, M.S. Glycosaminoglycans Affect the Interaction of Human Plasma Kallikrein with Plasminogen, Factor XII and Inhibitors. Braz. J. Med. Biol. Res. 2003, 36, 1055–1059. [Google Scholar] [CrossRef] [PubMed]
  94. Burrowes, C.E.; Habal, F.M.; Movat, H.Z. The Inhibition of Human Plasma Kallikrein by Antithrombin III. Thromb. Res. 1975, 7, 175–183. [Google Scholar] [CrossRef]
  95. Highsmith, R.F.; Rosenberg, R.D. The Inhibition of Human Plasmin by Human Antithrombin-Heparin Cofactor. J. Biol. Chem. 1974, 249, 4335–4338. [Google Scholar] [CrossRef]
  96. Lawson, J.H.; Butenas, S.; Ribarik, N.; Mann, K.G. Complex-Dependent Inhibition of Factor VIIa by Antithrombin III and Heparin. J. Biol. Chem. 1993, 268, 767–770. [Google Scholar] [CrossRef]
  97. Hallgren, J.; Lindahl, S.; Pejler, G. Structural Requirements and Mechanism for Heparin-Dependent Activation and Tetramerization of Human BetaI- and BetaII-Tryptase. J. Mol. Biol. 2005, 345, 129–139. [Google Scholar] [CrossRef]
  98. José Barbosa Pereira, P.; Bergner, A.; Macedo-Ribeiro, S.; Huber, R.; Matschiner, G.; Fritz, H.; Sommerhoff, C.P.; Bode, W. Human Beta-Tryptase Is a Ring-like Tetramer with Active Sites Facing a Central Pore. Nature 1998, 392, 306–311. [Google Scholar] [CrossRef]
  99. Fajardo, I.; Pejler, G. Formation of Active Monomers from Tetrameric Human Beta-Tryptase. Biochem. J. 2003, 369, 603. [Google Scholar] [CrossRef]
  100. Lindstedt, L.; Lee, M.; Kovanen, P.T. Chymase Bound to Heparin Is Resistant to Its Natural Inhibitors and Capable of Proteolyzing High Density Lipoproteins in Aortic Intimal Fluid. Atherosclerosis 2001, 155, 87–97. [Google Scholar] [CrossRef]
  101. Walter, M.; Plotnick, M.; Schechter, N.M. Inhibition of Human Mast Cell Chymase by Secretory Leukocyte Proteinase Inhibitor: Enhancement of the Interaction by Heparin. Arch. Biochem. Biophys. 1996, 327, 81–88. [Google Scholar] [CrossRef]
  102. Ermolieff, J.; Boudier, C.; Laine, A.; Meyer, B.; Bieth, J.G. Heparin Protects Cathepsin G against Inhibition by Protein Proteinase Inhibitors. J. Biol. Chem. 1994, 269, 29502–29508. [Google Scholar] [CrossRef] [PubMed]
  103. Ermolieff, J.; Duranton, J.; Petitou, M.; Bieth, J.G. Heparin Accelerates the Inhibition of Cathepsin G by Mucus Proteinase Inhibitor: Potent Effect of O-Butyrylated Heparin. Biochem. J. 1998, 330 Pt 3, 1369–1374. [Google Scholar] [CrossRef]
  104. Cadene, M.; Boudier, C.; De Marcillac, G.D.; Bieth, J.G. Influence of Low Molecular Mass Heparin on the Kinetics of Neutrophil Elastase Inhibition by Mucus Proteinase Inhibitor. J. Biol. Chem. 1995, 270, 13204–13209. [Google Scholar] [CrossRef] [PubMed]
  105. Faller, B.; Bieth, J.G.; Mely, Y.; Gerard, D. Heparin-Induced Conformational Change and Activation of Mucus Proteinase Inhibitor. Biochemistry 1992, 31, 8285–8290. [Google Scholar] [CrossRef]
  106. Frommherz, K.J.; Faller, B.; Bieth, J.G. Heparin Strongly Decreases the Rate of Inhibition of Neutrophil Elastase by Alpha 1-Proteinase Inhibitor. J. Biol. Chem. 1991, 266, 15356–15362. [Google Scholar] [CrossRef] [PubMed]
  107. Ledoux, D.; Merciris, D.; Barritault, D.; Caruelle, J.P. Heparin-like Dextran Derivatives as Well as Glycosaminoglycans Inhibit the Enzymatic Activity of Human Cathepsin G. FEBS Lett. 2003, 537, 23–29. [Google Scholar] [CrossRef] [PubMed]
  108. Evangelista, V.; Piccardoni, P.; Maugeri, N.; De Gaetano, G.; Cerletti, C. Inhibition by Heparin of Platelet Activation Induced by Neutrophil-Derived Cathepsin G. Eur. J. Pharmacol. 1992, 216, 401–405. [Google Scholar] [CrossRef]
  109. Spencer, J.L.; Stone, P.J.; Nugent, M.A. New Insights into the Inhibition of Human Neutrophil Elastase by Heparin. Biochemistry 2006, 45, 9104–9120. [Google Scholar] [CrossRef]
  110. Baici, A.; Bradamante, P. Interaction between Human Leukocyte Elastase and Chondroitin Sulfate. Chem. Biol. Interact. 1984, 51, 1–11. [Google Scholar] [CrossRef]
  111. Kostoulas, G.; Hörler, D.; Naggi, A.; Casu, B.; Baici, A. Electrostatic Interactions between Human Leukocyte Elastase and Sulfated Glycosaminoglycans: Physiological Implications. Biol. Chem. 1997, 378, 1481–1489. [Google Scholar] [CrossRef]
  112. Schenker, P.; Baici, A. Paradoxical Interactions between Modifiers and Elastase-2. FEBS J. 2010, 277, 2486–2495. [Google Scholar] [CrossRef]
  113. Dewald, B.; Rindler-Ludwig, R.; Bretz, U.; Baggiolini, M. Subcellular Localization and Heterogeneity of Neutral Proteases in Neutrophilic Polymorphonuclear Leukocytes. J. Exp. Med. 1975, 141, 709–723. [Google Scholar] [CrossRef]
  114. Li, Z.; Yasuda, Y.; Li, W.; Bogyo, M.; Katz, N.; Gordon, R.E.; Fields, G.B.; Brömme, D. Regulation of Collagenase Activities of Human Cathepsins by Glycosaminoglycans. J. Biol. Chem. 2004, 279, 5470–5479. [Google Scholar] [CrossRef] [PubMed]
  115. Li, Z.; Hou, W.S.; Escalante-Torres, C.R.; Gelb, B.D.; Bromme, D. Collagenase Activity of Cathepsin K Depends on Complex Formation with Chondroitin Sulfate. J. Biol. Chem. 2002, 277, 28669–28676. [Google Scholar] [CrossRef] [PubMed]
  116. Cherney, M.M.; Lecaille, F.; Kienitz, M.; Nallaseth, F.S.; Li, Z.; James, M.N.G.; Brömme, D. Structure-Activity Analysis of Cathepsin K/Chondroitin 4-Sulfate Interactions. J. Biol. Chem. 2011, 286, 8988–8998. [Google Scholar] [CrossRef]
  117. Aguda, A.H.; Panwar, P.; Du, X.; Nguyen, N.T.; Brayer, G.D.; Brömme, D. Structural Basis of Collagen Fiber Degradation by Cathepsin K. Proc. Natl. Acad. Sci. USA 2014, 111, 17474–17479. [Google Scholar] [CrossRef]
  118. Li, Z.; Kienetz, M.; Cherney, M.M.; James, M.N.G.; Brömme, D. The Crystal and Molecular Structures of a Cathepsin K:Chondroitin Sulfate Complex. J. Mol. Biol. 2008, 383, 78–91. [Google Scholar] [CrossRef] [PubMed]
  119. Li, Z.; Hou, W.S.; Brömme, D. Collagenolytic Activity of Cathepsin K Is Specifically Modulated by Cartilage-Resident Chondroitin Sulfates. Biochemistry 2000, 39, 529–536. [Google Scholar] [CrossRef] [PubMed]
  120. Novinec, M.; Kovačič, L.; Lenarĉič, B.; Baici, A. Conformational Flexibility and Allosteric Regulation of Cathepsin K. Biochem. J. 2010, 429, 379–389. [Google Scholar] [CrossRef]
  121. Sage, J.; Mallèvre, F.; Barbarin-Costes, F.; Samsonov, S.A.; Gehrcke, J.P.; Pisabarro, M.T.; Perrier, E.; Schnebert, S.; Roget, A.; Livache, T.; et al. Binding of Chondroitin 4-Sulfate to Cathepsin S Regulates Its Enzymatic Activity. Biochemistry 2013, 52, 6487–6498. [Google Scholar] [CrossRef] [PubMed]
  122. Denamur, S.; Chazeirat, T.; Maszota-Zieleniak, M.; Vivès, R.R.; Saidi, A.; Zhang, F.; Linhardt, R.J.; Labarthe, F.; Samsonov, S.A.; Lalmanach, G.; et al. Binding of Heparan Sulfate to Human Cystatin C Modulates Inhibition of Cathepsin L: Putative Consequences in Mucopolysaccharidosis. Carbohydr. Polym. 2022, 293, 119734. [Google Scholar] [CrossRef] [PubMed]
  123. Higgins, W.J.; Fox, D.M.; Kowalski, P.S.; Nielsen, J.E.; Worrall, D.M. Heparin Enhances Serpin Inhibition of the Cysteine Protease Cathepsin L. J. Biol. Chem. 2010, 285, 3722–3729. [Google Scholar] [CrossRef] [PubMed]
  124. Lemaire, P.A.; Huang, L.; Zhuo, Y.; Lu, J.; Bahnck, C.; Stachel, S.J.; Carroll, S.S.; Duong, L.T. Chondroitin Sulfate Promotes Activation of Cathepsin K. J. Biol. Chem. 2014, 289, 21562–21572. [Google Scholar] [CrossRef]
  125. Caglič, D.; Pungerčar, J.R.; Pejler, G.; Turk, V.; Turk, B. Glycosaminoglycans Facilitate Procathepsin B Activation through Disruption of Propeptide-Mature Enzyme Interactions. J. Biol. Chem. 2007, 282, 33076–33085. [Google Scholar] [CrossRef] [PubMed]
  126. Vasiljeva, O.; Dolinar, M.; Pungerčar, J.R.; Turk, V.; Turk, B. Recombinant Human Procathepsin S Is Capable of Autocatalytic Processing at Neutral PH in the Presence of Glycosaminoglycans. FEBS Lett. 2005, 579, 1285–1290. [Google Scholar] [CrossRef] [PubMed]
  127. Echtermeyer, F.; Bertrand, J.; Dreier, R.; Meinecke, I.; Neugebauer, K.; Fuerst, M.; Lee, Y.J.; Song, Y.W.; Herzog, C.; Theilmeier, G.; et al. Syndecan-4 Regulates ADAMTS-5 Activation and Cartilage Breakdown in Osteoarthritis. Nat. Med. 2009, 15, 1072–1076. [Google Scholar] [CrossRef]
  128. Crabbe, T.; Ioannou, C.; Docherty, A.J.P. Human Progelatinase A Can Be Activated by Autolysis at a Rate That Is Concentration-Dependent and Enhanced by Heparin Bound to the C-Terminal Domain. Eur. J. Biochem. 1993, 218, 431–438. [Google Scholar] [CrossRef] [PubMed]
  129. Crabbe, T.; O’Connell, J.P.; Smith, B.J.; Docherty, A.J.P. Reciprocated Matrix Metalloproteinase Activation: A Process Performed by Interstitial Collagenase and Progelatinase A. Biochemistry 1994, 33, 14419–14425. [Google Scholar] [CrossRef]
  130. Gao, G.; Plaas, A.; Thompson, V.P.; Jin, S.; Zuo, F.; Sandy, J.D. ADAMTS4 (Aggrecanase-1) Activation on the Cell Surface Involves C-Terminal Cleavage by Glycosylphosphatidyl Inositol-Anchored Membrane Type 4-Matrix Metalloproteinase and Binding of the Activated Proteinase to Chondroitin Sulfate and Heparan Sulfate on Syndecan-1. J. Biol. Chem. 2004, 279, 10042–10051. [Google Scholar] [CrossRef]
  131. Sørensen, H.P.; Vivès, R.R.; Manetopoulos, C.; Albrechtsen, R.; Lydolph, M.C.; Jacobsen, J.; Couchman, J.R.; Wewer, U.M. Heparan Sulfate Regulates ADAM12 through a Molecular Switch Mechanism. J. Biol. Chem. 2008, 283, 31920–31932. [Google Scholar] [CrossRef]
  132. Pozzi, N.; Chen, Z.; Gohara, D.W.; Niu, W.; Heyduk, T.; Di Cera, E. Crystal Structure of Prothrombin Reveals Conformational Flexibility and Mechanism of Activation. J. Biol. Chem. 2013, 288, 22734–22744. [Google Scholar] [CrossRef] [PubMed]
  133. Stubbs, M.T.; Oschkinat, H.; Mayr, I.; Huber, R.; Angliker, H.; Stone, S.R.; Bode, W. The Interaction of Thrombin with Fibrinogen. A Structural Basis for Its Specificity. Eur. J. Biochem. 1992, 206, 187–195. [Google Scholar] [CrossRef] [PubMed]
  134. Pechik, I.; Madrazo, J.; Mosesson, M.W.; Hernandez, I.; Gilliland, G.L.; Medved, L. Crystal Structure of the Complex between Thrombin and the Central “E” Region of Fibrin. Proc. Natl. Acad. Sci. USA 2004, 101, 2718–2723. [Google Scholar] [CrossRef] [PubMed]
  135. Baglin, T.P.; Carrell, R.W.; Church, F.C.; Esmon, C.T.; Huntington, J.A. Crystal Structures of Native and Thrombin-Complexed Heparin Cofactor II Reveal a Multistep Allosteric Mechanism. Proc. Natl. Acad. Sci. USA 2002, 99, 11079–11084. [Google Scholar] [CrossRef]
  136. Yun, T.H.; Baglia, F.A.; Myles, T.; Navaneetham, D.; López, J.A.; Walsh, P.N.; Leung, L.L.K. Thrombin Activation of Factor XI on Activated Platelets Requires the Interaction of Factor XI and Platelet Glycoprotein Ib Alpha with Thrombin Anion-Binding Exosites I and II, Respectively. J. Biol. Chem. 2003, 278, 48112–48119. [Google Scholar] [CrossRef]
  137. Fuentes-Prior, P.; Iwanaga, Y.; Huber, R.; Pagila, R.; Rumennik, G.; Seto, M.; Morser, J.; Light, D.R.; Bode, W. Structural Basis for the Anticoagulant Activity of the Thrombin-Thrombomodulin Complex. Nature 2000, 404, 518–525. [Google Scholar] [CrossRef] [PubMed]
  138. Carter, W.J.; Cama, E.; Huntington, J.A. Crystal Structure of Thrombin Bound to Heparin. J. Biol. Chem. 2005, 280, 2745–2749. [Google Scholar] [CrossRef]
  139. Sheehan, J.P.; Sadler, J.E. Molecular Mapping of the Heparin-Binding Exosite of Thrombin. Proc. Natl. Acad. Sci. USA 1994, 91, 5518–5522. [Google Scholar] [CrossRef]
  140. Li, C.Q.; Vindigni, A.; Sadler, J.E.; Wardell, M.R. Platelet Glycoprotein Ib Alpha Binds to Thrombin Anion-Binding Exosite II Inducing Allosteric Changes in the Activity of Thrombin. J. Biol. Chem. 2001, 276, 6161–6168. [Google Scholar] [CrossRef]
  141. Naskis, M.C.; Fenton, J.W.; Maraganoren, J.M.; Olson, S.T.; Shafer, J.A. The COOH-Terminal Domain of Hirudin. J. Biol. Chem. 1990, 265, 13484–13489. [Google Scholar]
  142. Krstenansky, J.L.; Mao, S.J. Antithrombin Properties of C-Terminus of Hirudin Using Synthetic Unsulfated N Alpha-Acetyl-Hirudin45-65. FEBS Lett. 1987, 211, 10–16. [Google Scholar] [CrossRef]
  143. Arocas, V.; Zingali, R.B.; Guillin, M.C.; Bon, C.; Jandrot-Perrus, M. Bothrojaracin: A Potent Two-Site-Directed Thrombin Inhibitor. Biochemistry 1996, 35, 9083–9089. [Google Scholar] [CrossRef] [PubMed]
  144. Zingali, R.B.; Jandrot-Perrus, M.; Guillin, M.C.; Bon, C. Bothrojaracin, a New Thrombin Inhibitor Isolated from Bothrops Jararaca Venom: Characterization and Mechanism of Thrombin Inhibition. Biochemistry 1993, 32, 10794–10802. [Google Scholar] [CrossRef] [PubMed]
  145. AR, R. Identification of Basic Residues in the Heparin-Binding Exosite of Factor Xa Critical for Heparin and Factor Va Binding. J. Biol. Chem. 2000, 275, 3320–3327. [Google Scholar] [CrossRef]
  146. Yang, L.; Manithody, C.; Rezaie, A.R. Localization of the Heparin Binding Exosite of Factor IXa. J. Biol. Chem. 2002, 277, 50756–50760. [Google Scholar] [CrossRef] [PubMed]
  147. Banner, D.W.; D’Arcy, A.; Chène, C.; Winkler, F.K.; Guha, A.; Konigsberg, W.H.; Nemerson, Y.; Kirchhofer, D. The Crystal Structure of the Complex of Blood Coagulation Factor VIIa with Soluble Tissue Factor. Nature 1996, 380, 41–46. [Google Scholar] [CrossRef]
  148. Francischetti, I.M.B.; Valenzuela, J.G.; Andersen, J.F.; Mather, T.N.; Ribeiro, J.M.C. Ixolaris, a Novel Recombinant Tissue Factor Pathway Inhibitor (TFPI) from the Salivary Gland of the Tick, Ixodes Scapularis: Identification of Factor X and Factor Xa as Scaffolds for the Inhibition of Factor VIIa/Tissue Factor Complex. Blood 2002, 99, 3602–3612. [Google Scholar] [CrossRef]
  149. Du, X.; Chen, N.L.H.; Wong, A.; Craik, C.S.; Brömme, D. Elastin Degradation by Cathepsin V Requires Two Exosites. J. Biol. Chem. 2013, 288, 34871–34881. [Google Scholar] [CrossRef]
  150. Sharma, V.; Panwar, P.; O’Donoghue, A.J.; Cui, H.; Guido, R.V.C.; Craik, C.S.; Brömme, D. Structural Requirements for the Collagenase and Elastase Activity of Cathepsin K and Its Selective Inhibition by an Exosite Inhibitor. Biochem. J. 2015, 465, 163–173. [Google Scholar] [CrossRef]
  151. Panwar, P.; Law, S.; Jamroz, A.; Azizi, P.; Zhang, D.; Ciufolini, M.; Brömme, D. Tanshinones That Selectively Block the Collagenase Activity of Cathepsin K Provide a Novel Class of Ectosteric Antiresorptive Agents for Bone. Br. J. Pharmacol. 2018, 175, 902. [Google Scholar] [CrossRef] [PubMed]
  152. Panwar, P.; Xue, L.; Søe, K.; Srivastava, K.; Law, S.; Delaisse, J.M.; Brömme, D. An Ectosteric Inhibitor of Cathepsin K Inhibits Bone Resorption in Ovariectomized Mice. J. Bone Miner. Res. 2017, 32, 2415–2430. [Google Scholar] [CrossRef] [PubMed]
  153. Robichaud, T.K.; Steffensen, B.; Fieldss, G.B. Exosite Interactions Impact Matrix Metalloproteinase Collagen Specificities. J. Biol. Chem. 2011, 286, 37535–37542. [Google Scholar] [CrossRef] [PubMed]
  154. Gao, W.; Anderson, P.J.; Majerus, E.M.; Tuley, E.A.; Sadler, J.E. Exosite Interactions Contribute to Tension-Induced Cleavage of von Willebrand Factor by the Antithrombotic ADAMTS13 Metalloprotease. Proc. Natl. Acad. Sci. USA 2006, 103, 19099–19104. [Google Scholar] [CrossRef]
  155. Santamaria, S.; Yamamoto, K.; Teraz-Orosz, A.; Koch, C.; Apte, S.S.; de Groot, R.; Lane, D.A.; Ahnström, J. Exosites in Hypervariable Loops of ADAMTS Spacer Domains Control Substrate Recognition and Proteolysis. Sci. Rep. 2019, 9, 10914. [Google Scholar] [CrossRef]
  156. Knapinska, A.M.; Dreymuller, D.; Ludwig, A.; Smith, L.; Golubkov, V.; Sohail, A.; Fridman, R.; Giulianotti, M.; LaVoi, T.M.; Houghten, R.A.; et al. SAR Studies of Exosite-Binding Substrate-Selective Inhibitors of A Disintegrin And Metalloprotease 17 (ADAM17) and Application as Selective in Vitro Probes. J. Med. Chem. 2015, 58, 5808. [Google Scholar] [CrossRef]
  157. Strisovsky, K.; Sharpe, H.J.; Freeman, M. Sequence-Specific Intramembrane Proteolysis: Identification of a Recognition Motif in Rhomboid Substrates. Mol. Cell 2009, 36, 1048. [Google Scholar] [CrossRef] [PubMed]
  158. Woskowicz, A.M.; Weaver, S.A.; Shitomi, Y.; Ito, N.; Itoh, Y. MT-LOOP-Dependent Localization of Membrane Type I Matrix Metalloproteinase (MT1-MMP) to the Cell Adhesion Complexes Promotes Cancer Cell Invasion. J. Biol. Chem. 2013, 288, 35149–35158. [Google Scholar] [CrossRef]
  159. Arnold, L.H.; Butt, L.E.; Prior, S.H.; Read, C.M.; Fields, G.B.; Pickford, A.R. The Interface between Catalytic and Hemopexin Domains in Matrix Metalloproteinase-1 Conceals a Collagen Binding Exosite. J. Biol. Chem. 2011, 286, 45073–45082. [Google Scholar] [CrossRef]
  160. Mehana, E.S.E.; Khafaga, A.F.; El-Blehi, S.S. The Role of Matrix Metalloproteinases in Osteoarthritis Pathogenesis: An Updated Review. Life Sci. 2019, 234, 116786. [Google Scholar] [CrossRef]
  161. Gresele, P.; Falcinelli, E.; Sebastiano, M.; Momi, S. Matrix Metalloproteinases and Platelet Function. Prog. Mol. Biol. Transl. Sci. 2017, 147, 133–165. [Google Scholar] [CrossRef] [PubMed]
  162. Mead, T.J.; Apte, S.S. ADAMTS Proteins in Human Disorders. Matrix Biol. 2018, 71–72, 225–239. [Google Scholar] [CrossRef] [PubMed]
  163. Ma, Z.; Mao, C.; Jia, Y.; Fu, Y.; Kong, W. Tissue Remodeling: From Regeneration to Fibrosis: Extracellular Matrix Dynamics in Vascular Remodeling. Am. J. Physiol. Cell Physiol. 2020, 319, C481. [Google Scholar] [CrossRef] [PubMed]
  164. Cui, N.; Hu, M.; Khalil, R.A. Biochemical and Biological Attributes of Matrix Metalloproteinases. Prog. Mol. Biol. Transl. Sci. 2017, 147, 1–73. [Google Scholar] [CrossRef] [PubMed]
  165. Santamaria, S.; Yamamoto, K.; Botkjaer, K.; Tape, C.; Dyson, M.R.; McCafferty, J.; Murphy, G.; Nagase, H. Antibody-Based Exosite Inhibitors of ADAMTS-5 (Aggrecanase-2). Biochem. J. 2015, 471, 391. [Google Scholar] [CrossRef] [PubMed]
  166. Santamaria, S.; Cuffaro, D.; Nuti, E.; Ciccone, L.; Tuccinardi, T.; Liva, F.; D’Andrea, F.; de Groot, R.; Rossello, A.; Ahnström, J. Exosite Inhibition of ADAMTS-5 by a Glycoconjugated Arylsulfonamide. Sci. Rep. 2021, 11, 949. [Google Scholar] [CrossRef] [PubMed]
  167. Aoki, K. Aptamer for ADAMTS5 and Use for Aptamer for ADAMTS5. U.S. Patent No. 11,473,089, 18 October 2022. [Google Scholar]
  168. Neuhold, L.A.; Killar, L.; Zhao, W.; Sung, M.L.A.; Warner, L.; Kulik, J.; Turner, J.; Wu, W.; Billinghurst, C.; Meijers, T.; et al. Postnatal Expression in Hyaline Cartilage of Constitutively Active Human Collagenase-3 (MMP-13) Induces Osteoarthritis in Mice. J. Clin. Investig. 2001, 107, 35–44. [Google Scholar] [CrossRef]
  169. Spicer, T.P.; Jiang, J.; Taylor, A.B.; Choi, J.Y.; Hart, P.J.; Roush, W.R.; Fields, G.B.; Hodder, P.S.; Minond, D. Characterization of Selective Exosite-Binding Inhibitors of Matrix Metalloproteinase 13 That Prevent Articular Cartilage Degradation in Vitro. J. Med. Chem. 2014, 57, 9598–9611. [Google Scholar] [CrossRef]
  170. Engel, C.K.; Pirard, B.; Schimanski, S.; Kirsch, R.; Habermann, J.; Klingler, O.; Schlotte, V.; Weithmann, K.U.; Wendt, K.U. Structural Basis for the Highly Selective Inhibition of MMP-13. Chem. Biol. 2005, 12, 181–189. [Google Scholar] [CrossRef]
  171. Johnson, A.R.; Pavlovsky, A.G.; Ortwine, D.F.; Prior, F.; Man, C.F.; Bornemeier, D.A.; Banotai, C.A.; Mueller, W.T.; McConnell, P.; Yan, C.; et al. Discovery and Characterization of a Novel Inhibitor of Matrix Metalloprotease-13 That Reduces Cartilage Damage in Vivo without Joint Fibroplasia Side Effects. J. Biol. Chem. 2007, 282, 27781–27791. [Google Scholar] [CrossRef]
  172. Piecha, D.; Weik, J.; Kheil, H.; Becher, G.; Timmermann, A.; Jaworski, A.; Burger, M.; Hofmann, M.W. Novel Selective MMP-13 Inhibitors Reduce Collagen Degradation in Bovine Articular and Human Osteoarthritis Cartilage Explants. Inflamm. Res. 2010, 59, 379–389. [Google Scholar] [CrossRef]
  173. Baragi, V.M.; Becher, G.; Bendele, A.M.; Biesinger, R.; Bluhm, H.; Boer, J.; Deng, H.; Dodd, R.; Essers, M.; Feuerstein, T.; et al. A New Class of Potent Matrix Metalloproteinase 13 Inhibitors for Potential Treatment of Osteoarthritis: Evidence of Histologic and Clinical Efficacy without Musculoskeletal Toxicity in Rat Models. Arthritis Rheum. 2009, 60, 2008–2018. [Google Scholar] [CrossRef] [PubMed]
  174. Udi, Y.; Grossman, M.; Solomonov, I.; Dym, O.; Rozenberg, H.; Moreno, V.; Cuniasse, P.; Dive, V.; Arroyo, A.G.; Sagi, I. Inhibition Mechanism of Membrane Metalloprotease by an Exosite-Swiveling Conformational Antibody. Structure 2015, 23, 104–115. [Google Scholar] [CrossRef] [PubMed]
  175. Ingvarsen, S.; Madsen, D.H.; Hillig, T.; Lund, L.R.; Holmbeck, K.; Behrendt, N.; Engelholm, L.H. Dimerization of Endogenous MT1-MMP Is a Regulatory Step in the Activation of the 72-KDa Gelatinase MMP-2 on Fibroblasts and Fibrosarcoma Cells. Biol. Chem. 2008, 389, 943–953. [Google Scholar] [CrossRef] [PubMed]
  176. Basu, B.; Correa De Sampaio, P.; Mohammed, H.; Fogarasi, M.; Corrie, P.; Watkins, N.A.; Smethurst, P.A.; English, W.R.; Ouwehand, W.H.; Murphy, G. Inhibition of MT1-MMP Activity Using Functional Antibody Fragments Selected against Its Hemopexin Domain. Int. J. Biochem. Cell Biol. 2012, 44, 393–403. [Google Scholar] [CrossRef] [PubMed]
  177. Xu, X.; Chen, Z.; Wang, Y.; Bonewald, L.; Steffensen, B. Inhibition of MMP-2 Gelatinolysis by Targeting Exodomain–Substrate Interactions. Biochem. J. 2007, 406, 147. [Google Scholar] [CrossRef] [PubMed]
  178. Santamaria, S.; de Groot, R. Monoclonal Antibodies against Metzincin Targets. Br. J. Pharmacol. 2019, 176, 52–66. [Google Scholar] [CrossRef] [PubMed]
  179. Cuffaro, D.; Ciccone, L.; Rossello, A.; Nuti, E.; Santamaria, S. Targeting Aggrecanases for Osteoarthritis Therapy: From Zinc Chelation to Exosite Inhibition. J. Med. Chem. 2022, 65, 13505–13532. [Google Scholar] [CrossRef]
  180. Li, K.; Tay, F.R.; Yiu, C.K.Y. The Past, Present and Future Perspectives of Matrix Metalloproteinase Inhibitors. Pharmacol. Ther. 2020, 207, 107465. [Google Scholar] [CrossRef]
  181. Blasi, J.; Chapman, E.R.; Link, E.; Binz, T.; Yamasaki, S.T.C.; Niemann, H.; Jahn, R. Botulinum Neurotoxin A Selectively Cleaves the Synaptic Protein SNAP-25. Nature 1993, 365, 160–163. [Google Scholar] [CrossRef]
  182. Xue, S.; Javor, S.; Hixon, M.S.; Janda, K.D. Probing BoNT/A Protease Exosites: Implications for Inhibitor Design and Light Chain Longevity. Biochemistry 2014, 53, 6820–6824. [Google Scholar] [CrossRef]
  183. Hu, X.; Legler, P.M.; Southall, N.; Maloney, D.J.; Simeonov, A.; Jadhav, A. Structural Insight into Exosite Binding and Discovery of Novel Exosite Inhibitors of Botulinum Neurotoxin Serotype A through in Silico Screening. J. Comput. Aided Mol. Des. 2014, 28, 765–778. [Google Scholar] [CrossRef]
  184. Xue, S.; Seki, H.; Remes, M.; Šilhár, P.; Janda, K. Examination of α-Exosite Inhibitors against Botulinum Neurotoxin A Protease through Structure-Activity Relationship Studies of Chicoric Acid. Bioorg Med. Chem. Lett. 2017, 27, 4956. [Google Scholar] [CrossRef] [PubMed]
  185. Bremer, P.T.; Xue, S.; Janda, K.D. Picolinic Acids as β-Exosite Inhibitors of Botulinum Neurotoxin A Light Chain. Chem. Commun. 2016, 52, 12521–12524. [Google Scholar] [CrossRef]
  186. Dong, J.; Thompson, A.A.; Fan, Y.; Lou, J.; Conrad, F.; Ho, M.; Pires-Alves, M.; Wilson, B.A.; Stevens, R.C.; Marks, J.D. A Single-Domain Llama Antibody Potently Inhibits the Enzymatic Activity of Botulinum Neurotoxin by Binding to the Non-Catalytic Alpha-Exosite Binding Region. J. Mol. Biol. 2010, 397, 1106–1118. [Google Scholar] [CrossRef] [PubMed]
  187. Fan, Y.; Geren, I.N.; Dong, J.; Lou, J.; Wen, W.; Conrad, F.; Smith, T.J.; Smith, L.A.; Ho, M.; Pires-Alves, M.; et al. Monoclonal Antibodies Targeting the Alpha-Exosite of Botulinum Neurotoxin Serotype/A Inhibit Catalytic Activity. PLoS ONE 2015, 10, e0135306. [Google Scholar] [CrossRef] [PubMed]
  188. Bannwarth, L.; Goldberg, A.B.; Chen, C.; Turk, B.E. Identification of Exosite-Targeting Inhibitors of Anthrax Lethal Factor by High Throughput Screening. Chem. Biol. 2012, 19, 875. [Google Scholar] [CrossRef] [PubMed]
  189. Monod, J.; Wyman, J.; Changeux, J.P. On the Nature of Allosteric Transitions: A Plausible Model. J. Mol. Biol. 1965, 12, 88–118. [Google Scholar] [CrossRef] [PubMed]
  190. Koshland, D.E.; Nemethy, J.G.; Filmer, D. Comparison of Experimental Binding Data and Theoretical Models in Proteins Containing Subunits. Biochemistry 1966, 5, 365–385. [Google Scholar] [CrossRef]
  191. Nussinov, R.; Tsai, C.J. Allostery in Disease and in Drug Discovery. Cell 2013, 153, 293–305. [Google Scholar] [CrossRef]
  192. Esmon, N.L.; Owen, W.G.; Esmon, C.T. Isolation of a Membrane-Bound Cofactor for Thrombin-Catalyzed Activation of Protein C. J. Biol. Chem. 1982, 257, 859–864. [Google Scholar] [CrossRef] [PubMed]
  193. Wells, C.M.; Di Cera, E. Thrombin Is a Sodium Ion Activated Enzyme. Biochemistry 1992, 31, 11721–11730. [Google Scholar] [CrossRef] [PubMed]
  194. Dang, O.D.; Vindigni, A.; Di Cera, E. An Allosteric Switch Controls the Procoagulant and Anticoagulant Activities of Thrombin. Proc. Natl. Acad. Sci. USA 1995, 92, 5977–5981. [Google Scholar] [CrossRef] [PubMed]
  195. Orthner, C.L.; Kosow, D.P. Evidence That Human α-Thrombin Is a Monovalent Cation-Activated Enzyme. Arch. Biochem. Biophys. 1980, 202, 63–75. [Google Scholar] [CrossRef] [PubMed]
  196. Ayala, Y.; Cera, E. Di Molecular Recognition by Thrombin. Role of the Slows→Fast Transition, Site-Specific Ion Binding Energetics and Thermodynamic Mapping of Structural Components. J. Mol. Biol. 1994, 235, 733–746. [Google Scholar] [CrossRef] [PubMed]
  197. Huntington, J.A. How Na+ Activates Thrombin—A Review of the Functional and Structural Data. Biol. Chem. 2008, 389, 1025–1035. [Google Scholar] [CrossRef] [PubMed]
  198. Bode, W.; Huber, R. Induction of the Bovine Trypsinogen-Trypsin Transition by Peptides Sequentially Similar to the N-Terminus of Trypsin. FEBS Lett. 1976, 68, 231–236. [Google Scholar] [CrossRef] [PubMed]
  199. Paul, S.; Castelliwo, F.J. Activation of Human Plasminogen by Equimolar Levels of Streptokinase. J. Biol. Chem. 1977, 252, 492–498. [Google Scholar] [CrossRef]
  200. Hemker, H.C.; Bas, B.M.; Muller, A.D. Activation of a Pro-Enzyme by a Stoichiometric Reaction with Another Protein: The Reaction between Prothrombin and Staphylocoagulase. Biochim. Biophys. Acta (BBA)-Protein Struct. 1975, 379, 180–188. [Google Scholar] [CrossRef]
  201. Friedrich, R.; Panizzi, P.; Fuentes-Prior, P.; Richter, K.; Verhamme, I.; Anderson, P.J.; Kawabata, S.I.; Huber, R.; Bode, W.; Bock, P.E. Staphylocoagulase Is a Prototype for the Mechanism of Cofactor-Induced Zymogen Activation. Nature 2003, 425, 535–539. [Google Scholar] [CrossRef]
  202. Novinec, M.; Korenč, M.; Caflisch, A.; Ranganathan, R.; Lenarčič, B.; Baici, A. A Novel Allosteric Mechanism in the Cysteine Peptidase Cathepsin K Discovered by Computational Methods. Nat. Commun. 2014, 5, 3287. [Google Scholar] [CrossRef]
  203. Novinec, M.; Lenarčič, B.; Baici, A. Probing the Activity Modification Space of the Cysteine Peptidase Cathepsin K with Novel Allosteric Modifiers. PLoS ONE 2014, 9, e106642. [Google Scholar] [CrossRef]
  204. Novinec, M. Computational Investigation of Conformational Variability and Allostery in Cathepsin K and Other Related Peptidases. PLoS ONE 2017, 12, e0182387. [Google Scholar] [CrossRef]
  205. Shiozaki, E.N.; Chai, J.; Rigotti, D.J.; Riedl, S.J.; Li, P.; Srinivasula, S.M.; Alnemri, E.S.; Fairman, R.; Shi, Y. Mechanism of XIAP-Mediated Inhibition of Caspase-9. Mol. Cell 2003, 11, 519–527. [Google Scholar] [CrossRef]
  206. Datta, D.; Scheer, J.M.; Romanowski, M.J.; Wells, J.A. An Allosteric Circuit in Caspase-1. J. Mol. Biol. 2008, 381, 1157–1167. [Google Scholar] [CrossRef]
  207. Schweizer, A.; Roschitzki-Voser, H.; Amstutz, P.; Briand, C.; Gulotti-Georgieva, M.; Prenosil, E.; Binz, H.K.; Capitani, G.; Baici, A.; Plückthun, A.; et al. Inhibition of Caspase-2 by a Designed Ankyrin Repeat Protein: Specificity, Structure, and Inhibition Mechanism. Structure 2007, 15, 625–636. [Google Scholar] [CrossRef] [PubMed]
  208. Hardy, J.A.; Lam, J.; Nguyen, J.T.; O’Brien, T.; Wells, J.A. Discovery of an Allosteric Site in the Caspases. Proc. Natl. Acad. Sci. USA 2004, 101, 12461. [Google Scholar] [CrossRef] [PubMed]
  209. Feldman, T.; Kabaleeswaran, V.; Jang, S.B.; Antczak, C.; Djaballah, H.; Wu, H.; Jiang, X. A Class of Allosteric, Caspase Inhibitors Identified by High-Throughput Screening. Mol. Cell 2012, 47, 585. [Google Scholar] [CrossRef] [PubMed]
  210. Flütsch, A.; Schroeder, T.; Barandun, J.; Ackermann, R.; Bühlmann, M.; Grütter, M.G.; Flütsch, A.; Schroeder, T.; Barandun, J.; Ackermann, R.; et al. Specific Targeting of Human Caspases Using Designed Ankyrin Repeat Proteins. Biol. Chem. 2014, 395, 1243–1252. [Google Scholar] [CrossRef]
  211. Van Es, J.H.; Van Gijn, M.E.; Riccio, O.; Van Den Born, M.; Vooijs, M.; Begthel, H.; Cozijnsen, M.; Robine, S.; Winton, D.J.; Radtke, F.; et al. Notch/Gamma-Secretase Inhibition Turns Proliferative Cells in Intestinal Crypts and Adenomas into Goblet Cells. Nature 2005, 435, 959–963. [Google Scholar] [CrossRef] [PubMed]
  212. Weggen, S.; Eriksen, J.L.; Das, P.; Sagi, S.A.; Wang, R.; Pietrzik, C.U.; Findlay, K.A.; Smith, T.E.; Murphy, M.P.; Bulter, T.; et al. A Subset of NSAIDs Lower Amyloidogenic Abeta42 Independently of Cyclooxygenase Activity. Nature 2001, 414, 212–216. [Google Scholar] [CrossRef] [PubMed]
  213. Mayer, S.C.; Kreft, A.F.; Harrison, B.; Abou-Gharbia, M.; Antane, M.; Aschmies, S.; Atchison, K.; Chlenov, M.; Cole, D.C.; Comery, T.; et al. Discovery of Begacestat, a Notch-1-Sparing Gamma-Secretase Inhibitor for the Treatment of Alzheimer’s Disease. J. Med. Chem. 2008, 51, 7348–7351. [Google Scholar] [CrossRef] [PubMed]
  214. Netzer, W.J.; Dou, F.; Cai, D.; Veach, D.; Jean, S.; Li, Y.; Bornmann, W.G.; Clarkson, B.; Xu, H.; Greengard, P. Gleevec Inhibits Beta-Amyloid Production but Not Notch Cleavage. Proc. Natl. Acad. Sci. USA 2003, 100, 12444–12449. [Google Scholar] [CrossRef] [PubMed]
  215. Shelton, C.C.; Zhu, L.; Chau, D.; Yang, L.; Wang, R.; Djaballah, H.; Zheng, H.; Li, Y.M. Modulation of γ-Secretase Specificity Using Small Molecule Allosteric Inhibitors. Proc. Natl. Acad. Sci. USA 2009, 106, 20228. [Google Scholar] [CrossRef] [PubMed]
  216. Campagna, J.; Vadivel, K.; Jagodzinska, B.; Jun, M.; Bilousova, T.; Spilman, P.; John, V. Evaluation of an Allosteric BACE Inhibitor Peptide to Identify Mimetics That Can Interact with the Loop F Region of the Enzyme and Prevent APP Cleavage. J. Mol. Biol. 2018, 430, 1566–1576. [Google Scholar] [CrossRef] [PubMed]
  217. Gutierrez, L.J.; Angelina, E.; Gyebrovszki, A.; Fülöp, L.; Peruchena, N.; Baldoni, H.A.; Penke, B.; Enriz, R.D. New Small-Size Peptides Modulators of the Exosite of BACE1 Obtained from a Structure-Based Design. J. Biomol. Struct. Dyn. 2017, 35, 413–426. [Google Scholar] [CrossRef]
  218. Ugbaja, S.C.; Lawal, I.A.; Abubakar, B.H.; Mushebenge, A.G.; Lawal, M.M.; Kumalo, H.M. Allostery Inhibition of BACE1 by Psychotic and Meroterpenoid Drugs in Alzheimer’s Disease Therapy. Molecules 2022, 27, 4372. [Google Scholar] [CrossRef] [PubMed]
  219. Atwal, J.K.; Chen, Y.; Chiu, C.; Mortensen, D.L.; Meilandt, W.J.; Liu, Y.; Heise, C.E.; Hoyte, K.; Luk, W.; Lu, Y.; et al. A Therapeutic Antibody Targeting BACE1 Inhibits Amyloid-β Production in Vivo. Sci. Transl. Med. 2011, 3, 84ra43. [Google Scholar] [CrossRef]
  220. Sela-Passwell, N.; Rosenblum, G.; Shoham, T.; Sagi, I. Structural and Functional Bases for Allosteric Control of MMP Activities: Can It Pave the Path for Selective Inhibition? Biochim. Biophys. Acta (BBA)-Mol. Cell Res. 2010, 1803, 29–38. [Google Scholar] [CrossRef]
  221. Kumar, L.; Nash, A.; Harms, C.; Planas-Iglesias, J.; Wright, D.; Klein-Seetharaman, J.; Sarkar, S.K. Allosteric Communications between Domains Modulate the Activity of Matrix Metalloprotease-1. Biophys. J. 2020, 119, 360–374. [Google Scholar] [CrossRef]
  222. Kamboj, S.; Harms, C.; Wright, D.; Nash, A.; Kumar, L.; Klein-Seetharaman, J.; Sarkar, S.K. Identification of Allosteric Fingerprints of Alpha-Synuclein Aggregates in Matrix Metalloprotease-1 and Substrate-Specific Virtual Screening with Single Molecule Insights. Sci. Rep. 2022, 12, 1–14. [Google Scholar] [CrossRef] [PubMed]
  223. Lubbe, L.; Sewell, B.T.; Woodward, J.D.; Sturrock, E.D. Cryo-EM Reveals Mechanisms of Angiotensin I-converting Enzyme Allostery and Dimerization. EMBO J. 2022, 41, 110550. [Google Scholar] [CrossRef] [PubMed]
  224. Mabanglo, M.F.; Houry, W.A. Recent Structural Insights into the Mechanism of ClpP Protease Regulation by AAA+ Chaperones and Small Molecules. J. Biol. Chem. 2022, 298, 101781. [Google Scholar] [CrossRef] [PubMed]
  225. Flynn, J.M.; Neher, S.B.; Kim, Y.I.; Sauer, R.T.; Baker, T.A. Proteomic Discovery of Cellular Substrates of the ClpXP Protease Reveals Five Classes of ClpX-Recognition Signals. Mol. Cell 2003, 11, 671–683. [Google Scholar] [CrossRef] [PubMed]
  226. El-Baba, T.J.; Lutomski, C.A.; Kantsadi, A.L.; Malla, T.R.; John, T.; Mikhailov, V.; Bolla, J.R.; Schofield, C.J.; Zitzmann, N.; Vakonakis, I.; et al. Allosteric Inhibition of the SARS-CoV-2 Main Protease: Insights from Mass Spectrometry Based Assays. Angew. Chem. Int. Ed. 2020, 59, 23544–23548. [Google Scholar] [CrossRef]
  227. Günther, S.; Reinke, P.Y.A.; Fernández-Garciá, Y.; Lieske, J.; Lane, T.J.; Ginn, H.M.; Koua, F.H.M.; Ehrt, C.; Ewert, W.; Oberthuer, D.; et al. X-Ray Screening Identifies Active Site and Allosteric Inhibitors of SARS-CoV-2 Main Protease. Science 2021, 372, 642–646. [Google Scholar] [CrossRef]
  228. Du, R.; Cooper, L.; Chen, Z.; Lee, H.; Rong, L.; Cui, Q. Discovery of Chebulagic Acid and Punicalagin as Novel Allosteric Inhibitors of SARS-CoV-2 3CLpro. Antivir. Res. 2021, 190, 105075. [Google Scholar] [CrossRef]
  229. Mirković, B.; Renko, M.; Turk, S.; Sosič, I.; Jevnikar, Z.; Obermajer, N.; Turk, D.; Gobec, S.; Kos, J. Novel Mechanism of CathepsinB Inhibition by Antibiotic Nitroxoline and Related Compounds. ChemMedChem 2011, 6, 1351–1356. [Google Scholar] [CrossRef]
  230. Sosič, I.; Mitrović, A.; Ćurić, H.; Knez, D.; Brodnik Žugelj, H.; Štefane, B.; Kos, J.; Gobec, S. Cathepsin B Inhibitors: Further Exploration of the Nitroxoline Core. Bioorg. Med. Chem. Lett. 2018, 28, 1239–1247. [Google Scholar] [CrossRef]
  231. Ulčakar, L.; Novinec, M. Inhibition of Human Cathepsins B and L by Caffeic Acid and Its Derivatives. Biomolecules 2021, 11, 31. [Google Scholar] [CrossRef]
  232. Goričan, T.; Ciber, L.; Petek, N.; Svete, J.; Novinec, M. Synthesis and Kinetic Characterization of Hyperbolic Inhibitors of Human Cathepsins K and S Based on a Succinimide Scaffold. Bioorg. Chem. 2021, 115, 105213. [Google Scholar] [CrossRef] [PubMed]
  233. Verespy, S.; Mehta, A.Y.; Afosah, D.; Al-Horani, R.A.; Desai, U.R. Allosteric Partial Inhibition of Monomeric Proteases. Sulfated Coumarins Induce Regulation, Not Just Inhibition, of Thrombin. Sci. Rep. 2016, 6, 24043. [Google Scholar] [CrossRef]
  234. Novinec, M.; Bembič, P.; Janković, M.; Kisilak, M.; Kljun, J.; Turel, I. Kinetic Characterization of Cerium and Gallium Ions as Inhibitors of Cysteine Cathepsins L, K, and S. Int. J. Mol. Sci. 2022, 23, 8993. [Google Scholar] [CrossRef] [PubMed]
  235. Kramer, L.; Renko, M.; Završnik, J.; Turk, D.; Seeger, M.A.; Vasiljeva, O.; Grütter, M.G.; Turk, V.; Turk, B. Non-Invasive in Vivo Imaging of Tumour-Associated Cathepsin B by a Highly Selective Inhibitory DARPin. Theranostics 2017, 7, 2806. [Google Scholar] [CrossRef] [PubMed]
  236. Yamamoto, K.; Murphy, G.; Troeberg, L. Extracellular Regulation of Metalloproteinases. Matrix Biol. 2015, 44–46, 255–263. [Google Scholar] [CrossRef] [PubMed]
  237. Dumin, J.A.; Dickeson, S.K.; Stricker, T.P.; Bhattacharyya-Pakrasi, M.; Roby, J.D.; Santoro, S.A.; Parks, W.C. Pro-Collagenase-1 (Matrix Metalloproteinase-1) Binds the A2β1 Integrin upon Release from Keratinocytes Migrating on Type I Collagen. J. Biol. Chem. 2001, 276, 29368–29374. [Google Scholar] [CrossRef]
  238. Piccard, H.; Van den Steen, P.E.; Opdenakker, G. Hemopexin Domains as Multifunctional Liganding Modules in Matrix Metalloproteinases and Other Proteins. J. Leukoc. Biol. 2007, 81, 870–892. [Google Scholar] [CrossRef]
  239. Lafleur, M.A.; Xu, D.; Hemler, M.E. Tetraspanin Proteins Regulate Membrane Type-1 Matrix Metalloproteinase-Dependent Pericellular Proteolysis. Mol. Biol. Cell 2009, 20, 2030–2040. [Google Scholar] [CrossRef]
  240. Takino, T.; Miyamori, H.; Kawaguchi, N.; Uekita, T.; Seiki, M.; Sato, H. Tetraspanin CD63 Promotes Targeting and Lysosomal Proteolysis of Membrane-Type 1 Matrix Metalloproteinase. Biochem. Biophys. Res. Commun. 2003, 304, 160–166. [Google Scholar] [CrossRef]
  241. Bharadwaj, A.G.; Holloway, R.W.; Miller, V.A.; Waisman, D.M. Plasmin and Plasminogen System in the Tumor Microenvironment: Implications for Cancer Diagnosis, Prognosis, and Therapy. Cancers 2021, 13, 1838. [Google Scholar] [CrossRef]
  242. Zhai, B.T.; Tian, H.; Sun, J.; Zou, J.B.; Zhang, X.F.; Cheng, J.X.; Shi, Y.J.; Fan, Y.; Guo, D.Y. Urokinase-Type Plasminogen Activator Receptor (UPAR) as a Therapeutic Target in Cancer. J. Transl. Med. 2022, 20, 135. [Google Scholar] [CrossRef]
  243. Leth, J.M.; Ploug, M. Targeting the Urokinase-Type Plasminogen Activator Receptor (UPAR) in Human Diseases With a View to Non-Invasive Imaging and Therapeutic Intervention. Front. Cell Dev. Biol. 2021, 9, 732015. [Google Scholar] [CrossRef]
  244. Turk, B.; Dolenc, I.; Žerovnik, E.; Turk, D.; Gubenšek, F.; Turk, V. Human Cathepsin B Is a Metastable Enzyme Stabilized by Specific Ionic Interactions Associated with the Active Site. Biochemistry 1994, 33, 14800–14806. [Google Scholar] [CrossRef] [PubMed]
  245. Aggarwal, N.; Sloane, B.F. Cathepsin B: Multiple Roles in Cancer. Proteom. Clin. Appl. 2014, 8, 427–437. [Google Scholar] [CrossRef] [PubMed]
  246. Cavallo-Medved, D.; Mai, J.; Dosescu, J.; Sameni, M.; Sloane, B.F. Caveolin-1 Mediates the Expression and Localization of Cathepsin B, pro-Urokinase Plasminogen Activator and Their Cell-Surface Receptors in Human Colorectal Carcinoma Cells. J. Cell Sci. 2005, 118, 1493–1503. [Google Scholar] [CrossRef] [PubMed]
  247. Cavallo-Medved, D.; Rudy, D.; Blum, G.; Bogyo, M.; Caglic, D.; Sloane, B.F. Live-Cell Imaging Demonstrates Extracellular Matrix Degradation in Association with Active Cathepsin B in Caveolae of Endothelial Cells during Tube Formation. Exp. Cell Res. 2009, 315, 1234–1246. [Google Scholar] [CrossRef] [PubMed]
  248. Mai, J.; Finley, R.L.; Waisman, D.M.; Sloane, B.F. Human Procathepsin B Interacts with the Annexin II Tetramer on the Surface of Tumor Cells. J. Biol. Chem. 2000, 275, 12806–12812. [Google Scholar] [CrossRef]
  249. Cui, Z.; Zeng, C.; Huang, F.; Yuan, F.; Yan, J.; Zhao, Y.; Zhou, Y.; Hankey, W.; Jin, V.X.; Huang, J.; et al. Cas13d Knockdown of Lung Protease Ctsl Prevents and Treats SARS-CoV-2 Infection. Nat. Chem. Biol. 2022, 18, 1056–1064. [Google Scholar] [CrossRef] [PubMed]
  250. Kameoka, J.; Tanaka, T.; Nojima, Y.; Schlossman, S.F.; Morimoto, C. Direct Association of Adenosine Deaminase with a T Cell Activation Antigen, CD26. Science 1993, 261, 466–469. [Google Scholar] [CrossRef]
  251. Ishii, T.; Ohnuma, K.; Murakami, A.; Takasawa, N.; Kobayashi, S.; Dang, N.H.; Schlossman, S.F.; Morimoto, C. CD26-Mediated Signaling for T Cell Activation Occurs in Lipid Rafts through Its Association with CD45RO. Proc. Natl. Acad. Sci. USA 2001, 98, 12138–12143. [Google Scholar] [CrossRef]
  252. Raj, V.S.; Mou, H.; Smits, S.L.; Dekkers, D.H.W.; Müller, M.A.; Dijkman, R.; Muth, D.; Demmers, J.A.A.; Zaki, A.; Fouchier, R.A.M.; et al. Dipeptidyl Peptidase 4 Is a Functional Receptor for the Emerging Human Coronavirus-EMC. Nature 2013, 495, 251–254. [Google Scholar] [CrossRef]
  253. Li, W.; Moore, M.J.; Vasllieva, N.; Sui, J.; Wong, S.K.; Berne, M.A.; Somasundaran, M.; Sullivan, J.L.; Luzuriaga, K.; Greeneugh, T.C.; et al. Angiotensin-Converting Enzyme 2 Is a Functional Receptor for the SARS Coronavirus. Nature 2003, 426, 450–454. [Google Scholar] [CrossRef] [PubMed]
  254. Zhou, P.; Yang, X.L.; Wang, X.G.; Hu, B.; Zhang, L.; Zhang, W.; Si, H.R.; Zhu, Y.; Li, B.; Huang, C.L.; et al. A Pneumonia Outbreak Associated with a New Coronavirus of Probable Bat Origin. Nature 2020, 579, 270–273. [Google Scholar] [CrossRef] [PubMed]
  255. Gupta, S.L.; Jaiswal, R.K. Neutralizing Antibody: A Savior in the Covid-19 Disease. Mol. Biol. Rep. 2022, 49, 2465–2474. [Google Scholar] [CrossRef] [PubMed]
  256. Zhang, H.; Lv, P.; Jiang, J.; Liu, Y.; Yan, R.; Shu, S.; Hu, B.; Xiao, H.; Cai, K.; Yuan, S.; et al. Advances in Developing ACE2 Derivatives against SARS-CoV-2. Lancet Microbe 2023, 4, e369–e378. [Google Scholar] [CrossRef] [PubMed]
  257. Jarocki, V.M.; Tacchi, J.L.; Djordjevic, S.P. Non-Proteolytic Functions of Microbial Proteases Increase Pathological Complexity. Proteomics 2015, 15, 1075–1088. [Google Scholar] [CrossRef] [PubMed]
  258. Hytönen, J.; Haataja, S.; Gerlach, D.; Podbielski, A.; Finne, J. The SpeB Virulence Factor of Streptococcus Pyogenes, a Multifunctional Secreted and Cell Surface Molecule with Strepadhesin, Laminin-Binding and Cysteine Protease Activity. Mol. Microbiol. 2001, 39, 512–519. [Google Scholar] [CrossRef] [PubMed]
  259. Beckmann, C.; Waggoner, J.D.; Harris, T.O.; Tamura, G.S.; Rubens, C.E. Identification of Novel Adhesins from Group B Streptococci by Use of Phage Display Reveals That C5a Peptidase Mediates Fibronectin Binding. Infect. Immun. 2002, 70, 2869–2876. [Google Scholar] [CrossRef] [PubMed]
  260. Rudzińska, M.; Daglioglu, C.; Savvateeva, L.V.; Kaci, F.N.; Antoine, R.; Zamyatnin, A.A. Current Status and Perspectives of Protease Inhibitors and Their Combination with Nanosized Drug Delivery Systems for Targeted Cancer Therapy. Drug Des. Dev. Ther. 2021, 15, 9–20. [Google Scholar] [CrossRef]
  261. Turk, B. Targeting Proteases: Successes, Failures and Future Prospects. Nat. Rev. Drug Dis. 2006, 5, 785–799. [Google Scholar] [CrossRef]
Figure 1. Mechanisms of regulation of proteolytic enzymes illustrated using the example of human thrombin. Orthosteric regulation, in which an effector is bound to the active site, is an established approach to modulate enzyme activity. However, our understanding goes beyond the conventional paradigm and also includes complicated mechanisms that go beyond the active site. In allosteric regulation, enzyme activity is modulated by binding to a site other than the active site, which changes the conformation of the enzyme. Regulation by exosites involves interactions with specific regions outside the active site that influence the activity and function of the enzyme. Zymogen activation refers to the process by which a peptidase precursor is converted to its active form through self-cleavage or other activation mechanisms. Taken together, these different regulatory mechanisms influence the activity and specificity of the enzyme.
Figure 1. Mechanisms of regulation of proteolytic enzymes illustrated using the example of human thrombin. Orthosteric regulation, in which an effector is bound to the active site, is an established approach to modulate enzyme activity. However, our understanding goes beyond the conventional paradigm and also includes complicated mechanisms that go beyond the active site. In allosteric regulation, enzyme activity is modulated by binding to a site other than the active site, which changes the conformation of the enzyme. Regulation by exosites involves interactions with specific regions outside the active site that influence the activity and function of the enzyme. Zymogen activation refers to the process by which a peptidase precursor is converted to its active form through self-cleavage or other activation mechanisms. Taken together, these different regulatory mechanisms influence the activity and specificity of the enzyme.
Ijms 24 17120 g001
Figure 2. Crystal structures of zymogens. (a) Superposition of the active sites of the serine peptidase trypsinogen (orange; PDB 1TGN) and trypsin (gray; PDB 1MCT). Residues His57, Asp102, and Ser195 form the active site. After cleavage of the activation hexapeptide, the newly formed N-terminus (Ile16) forms a salt bridge with Asp194, leading to a conformational change in the S1 binding pocket. (b) Structure of MMP-1 (PDB 1SU3), a collagenolytic member of the matrix metalloprotease family. It is a multidomain protein consisting of a propeptide (orange), a hemopexin domain (green), and a catalytic domain (blue). The free Cys92 residue in the propeptide (orange) of MMP-1 interacts as a fourth ligand for the zinc ion in the active site. The cysteine switch excludes water from the coordination sphere of zinc inactivating metalloproteinase.
Figure 2. Crystal structures of zymogens. (a) Superposition of the active sites of the serine peptidase trypsinogen (orange; PDB 1TGN) and trypsin (gray; PDB 1MCT). Residues His57, Asp102, and Ser195 form the active site. After cleavage of the activation hexapeptide, the newly formed N-terminus (Ile16) forms a salt bridge with Asp194, leading to a conformational change in the S1 binding pocket. (b) Structure of MMP-1 (PDB 1SU3), a collagenolytic member of the matrix metalloprotease family. It is a multidomain protein consisting of a propeptide (orange), a hemopexin domain (green), and a catalytic domain (blue). The free Cys92 residue in the propeptide (orange) of MMP-1 interacts as a fourth ligand for the zinc ion in the active site. The cysteine switch excludes water from the coordination sphere of zinc inactivating metalloproteinase.
Ijms 24 17120 g002
Figure 3. Inhibition of C5 zymogen activation by CirpT. The final stages of the complement activation pathway involve the activation of C5 by the peptidase C5 convertase. The convertase cleaves C5 into the proinflammatory fragments C5a and C5b. However, certain blood-sucking organisms have evolved a strategic mechanism to inhibit the activation of C5, effectively dampening the host immune response. One such inhibitor is CirpT1, which is secreted by the tick Rhipicephalus puchellus. (a) Binding of CirpT1 inhibitor (orange) to complement component C5 (PDB 6RQJ). Binding occurs via the β-chain of C5 (green). (b) Interaction between C5 and cobra venom factor (CVP; blue), which serves as an analog of the C3b fragment of C5 convertase (PDB 3PVM). The CVP or C5 convertase binding site coincides with the binding site of CirpT. This convergence of binding sites allows CirpT1 to effectively interfere with C5 activation [57].
Figure 3. Inhibition of C5 zymogen activation by CirpT. The final stages of the complement activation pathway involve the activation of C5 by the peptidase C5 convertase. The convertase cleaves C5 into the proinflammatory fragments C5a and C5b. However, certain blood-sucking organisms have evolved a strategic mechanism to inhibit the activation of C5, effectively dampening the host immune response. One such inhibitor is CirpT1, which is secreted by the tick Rhipicephalus puchellus. (a) Binding of CirpT1 inhibitor (orange) to complement component C5 (PDB 6RQJ). Binding occurs via the β-chain of C5 (green). (b) Interaction between C5 and cobra venom factor (CVP; blue), which serves as an analog of the C3b fragment of C5 convertase (PDB 3PVM). The CVP or C5 convertase binding site coincides with the binding site of CirpT. This convergence of binding sites allows CirpT1 to effectively interfere with C5 activation [57].
Ijms 24 17120 g003
Figure 4. Thrombin inhibition by the antithrombin–heparin complex. (a) Structure of human antithrombin in its inactive state (PDB 1E05). (b) Structure of active human antithrombin in complex with a pentasaccharide (orange), which resembles the binding of HP (PDB 1E03). This binding leads to conformational changes, including changes in the reactive center loop (RCL; yellow), which is critical for thrombin recognition and inhibition. The HP-induced conformational change increases the inhibitory effect of antithrombin. (c) Complex of activated antithrombin (green), HP (orange), and thrombin (blue). RCL (yellow) represents a larger portion of the interaction surface between both proteins.
Figure 4. Thrombin inhibition by the antithrombin–heparin complex. (a) Structure of human antithrombin in its inactive state (PDB 1E05). (b) Structure of active human antithrombin in complex with a pentasaccharide (orange), which resembles the binding of HP (PDB 1E03). This binding leads to conformational changes, including changes in the reactive center loop (RCL; yellow), which is critical for thrombin recognition and inhibition. The HP-induced conformational change increases the inhibitory effect of antithrombin. (c) Complex of activated antithrombin (green), HP (orange), and thrombin (blue). RCL (yellow) represents a larger portion of the interaction surface between both proteins.
Ijms 24 17120 g004
Figure 5. Crystal structure of thrombin. (a) Thrombin consists of four domains: the Gla domain, two kringle domains, and the trypsin-like catalytic domain. (b) Catalytic domain of thrombin, which includes two exosites in addition to the active site. Exosite I is located near the active site. Exosite II is located near the C-terminus. (c) Crystal structure of thrombin with bound huridin (magenta). Huridin is a natural inhibitor of thrombin isolated from the leech Hirudo meidicinalis. Huridin binds to the active site and exosite I of thrombin.
Figure 5. Crystal structure of thrombin. (a) Thrombin consists of four domains: the Gla domain, two kringle domains, and the trypsin-like catalytic domain. (b) Catalytic domain of thrombin, which includes two exosites in addition to the active site. Exosite I is located near the active site. Exosite II is located near the C-terminus. (c) Crystal structure of thrombin with bound huridin (magenta). Huridin is a natural inhibitor of thrombin isolated from the leech Hirudo meidicinalis. Huridin binds to the active site and exosite I of thrombin.
Ijms 24 17120 g005
Figure 6. Illustration of the conformational changes of thrombin. (a) Protein structures of the “slow” (gray; PDB 1RD3) and “fast” (green; PDB 1PPB) conformations of thrombin. In the “slow” conformation, the closure of the active site can be seen. The transition between conformations is Na+ dependent. (b) Active site of “slow” and “fast” thrombin. In the “slow” conformation, a rotation of residues Glu192 and Gly193 occurs, allowing the formation of hydrogen bonds (magenta) that undermine the structure of the oxyanion hole.
Figure 6. Illustration of the conformational changes of thrombin. (a) Protein structures of the “slow” (gray; PDB 1RD3) and “fast” (green; PDB 1PPB) conformations of thrombin. In the “slow” conformation, the closure of the active site can be seen. The transition between conformations is Na+ dependent. (b) Active site of “slow” and “fast” thrombin. In the “slow” conformation, a rotation of residues Glu192 and Gly193 occurs, allowing the formation of hydrogen bonds (magenta) that undermine the structure of the oxyanion hole.
Ijms 24 17120 g006
Figure 7. Crystal structures of human caspase-7 in complex with the allosteric inhibitors (a) DICA (PDB 1SHJ) and (b) NSC321205 (PDB 4FEA). Caspase-7 is shown in cartoon representation and the inhibitors are shown as spheres (orange). Each subunit of the dimer is colored in a different shade.
Figure 7. Crystal structures of human caspase-7 in complex with the allosteric inhibitors (a) DICA (PDB 1SHJ) and (b) NSC321205 (PDB 4FEA). Caspase-7 is shown in cartoon representation and the inhibitors are shown as spheres (orange). Each subunit of the dimer is colored in a different shade.
Ijms 24 17120 g007
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Obaha, A.; Novinec, M. Regulation of Peptidase Activity beyond the Active Site in Human Health and Disease. Int. J. Mol. Sci. 2023, 24, 17120. https://doi.org/10.3390/ijms242317120

AMA Style

Obaha A, Novinec M. Regulation of Peptidase Activity beyond the Active Site in Human Health and Disease. International Journal of Molecular Sciences. 2023; 24(23):17120. https://doi.org/10.3390/ijms242317120

Chicago/Turabian Style

Obaha, Ana, and Marko Novinec. 2023. "Regulation of Peptidase Activity beyond the Active Site in Human Health and Disease" International Journal of Molecular Sciences 24, no. 23: 17120. https://doi.org/10.3390/ijms242317120

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop